paper_id
stringlengths
9
16
version
stringclasses
26 values
yymm
stringclasses
311 values
created
timestamp[s]
title
stringlengths
6
335
secondary_subfield
sequencelengths
1
8
abstract
stringlengths
25
3.93k
primary_subfield
stringclasses
124 values
field
stringclasses
20 values
fulltext
stringlengths
0
2.84M
1609.02202
1
1609
2016-09-07T21:50:38
Deconstructing Odorant Identity via Primacy in Dual Networks
[ "q-bio.NC" ]
In the olfactory system, odor percepts retain their identity despite substantial variations in concentration, timing, and background. We propose a novel strategy for encoding intensity-invariant stimuli identity that is based on representing relative rather than absolute values of the stimulus features. Because, in this scheme, stimulus identity depends on relative amplitudes of stimulus features, identity becomes invariant with respect to variations in intensity and monotonous non-linearities of neuronal responses. In the olfactory system, stimulus identity can be represented by the identities of the p strongest responding odorant receptor types out of a species dependent complement. We show that this information is sufficient to recover sparse stimuli (odorants) via elastic net loss minimization. Such a minimization has to be performed under constraints imposed by the relationships between stimulus features. We map this problem onto the dual problem of minimizing a functional of Lagrange multipliers. The dual problem, in turn, can be solved by a neural network whose Lyapunov function represents the dual Lagrangian. We thus propose that networks in the piriform cortex compute odorant identity and implement dual computations with the sparse activities of individual neurons representing the Lagrange multipliers.
q-bio.NC
q-bio
Deconstructing Odorant Identity via Primacy in Dual Networks Daniel Kepple1, Hamza Giaffar1, Dmitry Rinberg2, and Alexei Koulakov1 1 Cold Spring Harbor Laboratory, Cold Spring Harbor, NY 2 NYU Neuroscience Institute, NYU Langone Medical center, New York, NY Abstract In the olfactory system, odor percepts retain their identity despite substantial variations in concentration, timing, and background. We propose a novel strategy for encoding intensity-invariant stimuli identity that is based on representing relative rather than absolute values of the stimulus features. Because, in this scheme, stimulus identity depends on relative amplitudes of stimulus features, identity becomes invariant with respect to variations in intensity and monotonous non-linearities of neuronal responses. In the olfactory system, stimulus identity can be represented by the identities of the p strongest responding odorant receptor types out of a species dependent complement. We show that this information is sufficient to recover sparse stimuli (odorants) via elastic net loss minimization. Such a minimization has to be performed under constraints imposed by the relationships between stimulus features. We map this problem onto the dual problem of minimizing a functional of Lagrange multipliers. The dual problem, in turn, can be solved by a neural network whose Lyapunov function represents the dual Lagrangian. We thus propose that networks in the piriform cortex compute odorant identity and implement dual computations with the sparse activities of individual neurons representing the Lagrange multipliers. 1 Introduction Sensory systems face the problem of computing stimulus identity invariant to several features. The olfactory system, for example, has to compute stimulus identity despite substantial variation in the absolute concentrations of molecules present in the stimulus. This computation is necessary to enable navigation in chemical gradients or within variable odorant plumes. How can the olfactory system robustly represent odorant identity despite variable stimulus intensity? The first step of olfactory processing involves odorants binding to and activating a set of molecular sensors known as olfactory receptors. Olfactory receptors are proteins expressed by olfactory sensory neurons (OSNs) located in the olfactory epithelium. Most mammalian olfactory systems contain ~1000 types olfactory receptors, while humans rely on the responses of only 350 [1; 2; 3]. Importantly, every OSN expresses only a single type of olfactory receptor chosen randomly out of the large ensemble. Odorant identity is therefore represented in the patterns of activation of olfactory receptor proteins and, by extension, OSNs. How the olfactory system represents odorant identity is not clearly understood. Here we examine the hypothesis that stimulus identity is inferred on the basis of the relative amplitudes of responses of OSNs. In particular, we propose that odorant identity can be defined by specifying the p strongest responding. We call this type of representation the primacy model. According to this primacy coding scheme, in which the olfactory system uses relative rather than absolute receptor responses, odorant representations become independent of absolute odorant concentration, leading to a concentration-invariant representation of odor identity. We formulate the identity decoding scheme using a dual Lagrange- 1 Karush-Kuhn-Tucker problem. We show that this dual problem can be solved by a neural network that we call a dual network. Dual networks that solve the primacy problem share many features with real olfactory networks. We therefore derive gross structure of olfactory circuits from first principles, based on the primacy model. 2 Dual Networks 2.1 Representing odorants by sparse vectors. 6 M ~ 10 . Each component of this vector Ethologically important odorant stimuli are mixtures of monomolecular components. Such stimuli can be represented by a vector of concentrations x jx is equal to the concentration of an individual monomolecular component numbered by the index j . The number of potential monomolecular components, which is equal to the dimensionality of vector x , will be denoted here by M . This number has been estimated to be on the order of several million, , based on the count of potentially volatile molecules lighter than 300 Daltons in the popular database PubChem [4]. However, odorant mixtures cannot contain all of these molecules at the same time and are necessarily represented by sparse vectors x . For the purposes of the olfactory system, the sparseness of the concentration vector is further increased by the inability of the system to detect or recognize individual components. Indeed, psychophysical studies suggest that human observers can detect roughly 12 monomolecular components of the vector x [5]. Therefore, we suggest that realistic odorants can be defined by concentration vectors x of high dimensionality ( Odorant mixtures are further represented by the responses of olfactory receptor neurons r responses can be approximated by linear nonlinear functions of vector x model of receptor with a single binding site and no cooperativity, the law of mass action yields (1) . These . Indeed, in the simplest ) and very few non-zero elements ( ~ 10 ~ 10 ). M K 6 r F y ( i i  ) y i   , A x ij j j i 1... N  Here index enumerates the olfactory receptor types. The total number of olfactory receptors N varies between 350 in humans and 1100 in rodents. Matrix ijA contains affinities of molecules of type j to the receptors of typei . is the nonlinear function describing the activation of a receptor. The problem solved by the olfactory system can be jx given the set of responses of olfactory sensory formulated as follows: find the odorant stimulus neurons ir . Because the responses iy via a simple monotonic function F , we can assume that networks analyzing responses of receptor neurons have the linear component of response iy available to them. Overall, the problem of olfactory decoding is this: find the sparse vector iy containing olfactory receptor responses. ir are related to their inputs jx given vector F y ( / (1   y y ) ) given y 2.2 Sparse olfactory stimulus recovery. The problem formulated above (find x ) can be reduced to solving the system of linear equations (1). This problem is not entirely trivial, because the number of unknowns (components of x ) is substantially larger than the number of equations (components of y M , N 3 ). Some relief arrives from the fact that the vector of unknowns is sparse. A sparse vector can be recovered from a system of linear equations using arguments from compressed sensing [6; 7; 8]. The vector x can be found exactly, despite containing more components than , ~ 10 ~ 10 6 2  j x j  y Ax  x 0  j  argmin , one can use the method of sparse signal recovery via exactly, the following certain condition has to be met relating equations. To determine x given y minimization of the 1l norm [6; 7; 8]:  x To be able to reconstruct x parameters M , N , and K : This condition is the necessary condition for sparse signal recovery using 1l norm minimization obtained by Donoho and Tanner [6; 7; 8]. For we obtain the following condition for the number of olfactory receptor types necessary for the recovery of the stimulus: N  ). Thus, equation (1) can in principle be solved with the existing parameters in the olfactory system. This means that the olfactory system can reconstruct monomolecular components given the responses of , which is satisfied for both humans ( olfactory receptors. ) and mice ( N K 1000 ~ 10 ~ 10 N  N  ~ 10 200 350 N  1000 and log M (2) (3) K M 6  K  results in the doubling of the vector of receptor responses y The solution (2) is limited in that it is not concentration invariant. Doubling the concentration vector x . Solving equation (2) under the constraint of doubled vector y in turn reconstructs the doubled vector of concentrations x . Our goal, however, is to build a representation that is concentration invariant. This means that we would like to obtain a framework in which doubling of the receptor responses y does not affect the reconstructed stimulus, i.e. a concentration-invariant odorant identity. We therefore propose a primacy-based decoding model that makes this possible.  Ax  , and the resulting receptor response vector y 1000 2.3 Concentration invariant decoding algorithm via primacy. To make odorant representation invariant to concentration, we propose to reformulate the sparse recovery problem. In its simplest form, for each input vector x  Ax , we propose to isolate the set of p strongest responding receptor neurons ( ) . We call this set the primacy set P . The complementary set P includes weaker responding receptor types. For these two sets we can state that y Here y means the set of components of vector y these constraints, the sparse recovery problem (2) can be reformulated as follows: P that belong to the primacy group P . With argmin p  (5) (4)  x  x y P   When seeking a minimum in this equation, we have to use the relationship y Ax , however, we do not use knowledge about exact values of the vector y , only the relationships between its components. This approach can be used to recover the concentration vector x , however, the jy . This amplitude of this vector does not depend on the absolute values of the neural responses is due to the non-parametric formulation of relative rather than absolute values of receptor responses in the sparse signal recovery problem (5). For example, doubling each component of vector y , and, thus, the reconstructed stimulus x is not affected. This decoding scheme therefore yields a reconstruction of the stimulus x that is concentration invariant: many x stimuli with the same direction and different length will yield a reconstructed concentration vector x that is exactly the same. We therefore suggest that the does not change the set of inequalities y y    j j y x y  P 0  j P 3  y  , thus leading to the same minimum 1l norm x primacy constraints result in recovery of concentration-invariant stimuli, leading to the decoding of the concentration-invariant odorant identity. Similarly, computing a monotonically increasing non-linear function of each component of y , such as that given by equation (1), does not affect the reconstructed vector x . Such an operation does not affect the constraint y . Within our approach the odorant identity is therefore invariant with respect to the non-linear monotonic transformations of receptor responses. One can therefore use linear and non-linear models of receptor responses i.e. y interchangingly, since this choice does not affect the solution. Overall, using relative rather than absolute values of sensor responses to reconstruct stimuli makes reconstruction invariant to a set of transformations, such as stimulus intensity etc. Although we demonstrated this idea for the particular example of primacy coding, we propose that this neural relativity rule could be used more generally by neural systems to transmit and recover signals invariant to various transformations. Instead of minimizing the 1l norm alone to find a sparse solution, one can minimize the elastic net functional:     and r arg min x  j x 2 j 2 (6)     x   P j y x y  P 0  j ~ y  that differ in their primacy code 0 . For The equation (6) is the same as the 1l norm minimization (5) when the parameter sufficiently small , equation (6) yields sparse solutions similar to equation (5). Problem (6), however allows a more straightforward formulation in the dual space. We will therefore use equation (6), i.e. elastic net minimization, to recover olfactory stimuli for the rest of the paper, with a sufficiently small . 2.4 Number of receptors needed to implement primacy coding. To estimate the number of primary receptors p needed to recover the stimulus we estimate a possible number of combinations in y . The estimate for the number of combinations in x number of combinations of y following condition: For a typical mammal, such as a mouse, for which the number of olfactory receptors is 1000 approximately . For humans, the number of functional olfactory N  and, therefore the primacy number (the number of primacy receptor types) receptors is required is somewhat higher, . In both of these estimates we assumed that the potential 6 . If one assumes number of types of molecules available in the environment is close to ~ 12 that the number of discernable molecular components in each mixture is , as follows which is substantially less from human psychophysics [5], we obtain the primacy number than the total number of receptors is. has to exceed the number of combinations in x . Assuming that for an exact reconstruction of x the , we arrive to the N  350 , we obtain x M ~ 10 K p K K / log NC p  ~p log 2.4 1000 N  (7) 30 M M N K N ~ p   p . K 2 p 2.5 Duality transformation. Since problem (6) includes various inequality constraints, one can use the Lagrange multiplier method to solve it. To present the problem in Lagrange form, one has to reformulate the conditions (4) in a more convenient form. Indeed, minimization of 1l norm or elastic net functional of x connected to y  Ax )  x  0 . To obtain a non-zero solution, one has to introduce a finite scale results in the solution into the conditions (4). A set of equivalent conditions with constrained magnitude is  via a set of positive coefficients A ( y   (8) ,  y y P P 4 j j ( i 1 y i (9) A x ij 0  iu   for i P and 0  is the scale parameter. By introducing a sign variable 1 where iu   for i P , equation (8) can be rewritten as a single set of conditions Here u y  ) i   with receptor i to odorant j affinities given by a set of non-negative ijA . Problem (6) combined with constraints (9) represents the primal optimization numbers problem. To transform the primal minimization problem (6) to the dual problem, we introduce the i, with Lagrangian with two sets of Lagrange multipliers, i N jx  0 correspond to the constraints (9), while the latter, 1.. for . The full Lagrangian for the elastic net minimization problem (6) is as follows  L x ) ( , ,   (10)   i j are supposed to have non-negative values, In this equation, the Lagrange multipliers j   [9]. To transform the primal problem into the dual problem, we minimize the i.e. Lagrangian (10) with respect to x if no constraints were present to compute the optimal value of x denoted here as x  . We then compute the value of the Lagrangian in the minimum    , ) , ) , (     :  L x ( , ,   j . The former multipliers, j , enforce the constraint  x     i and i and M  L x ( 1.. j u  i i A x ij   (11) / 2) 1) / x  0 )   j ) /       x ( x    , i x j 0  ( j 2 j j j j j  j  j (   i i   j u A i ij i  u A ij i 1 2    j  2 2 j (  j )  (12)  , )    (     G ik j Here   k u G i ik i  ik u k  1   ij (   u  1   A ij ) 1 2   i A A ij kj  i i in the Gramm matrix for rows of matrix A . According to optimization j  , )    and . These values can be used to find the solution of the primal problem using equation (11). The  is to be maximized to find the optimal values of  theory [9] the dual Lagrangian   dual problem can therefore be formulated as follows (13)  )     (      ) ( , ,  arg max (   j 0,   0 i , The reason why the dual problem (13) is preferable to the primal problem (6) lies in the simplicity j   are easier to implement than inequalities (9). The of the constraints. The constraints main observation that we make in this study is that, for the nervous system, it is especially easy to impose non-negativity constraints, because neural responses are described by firing rates that cannot fall below zero. Motivated by this observation, we will argue below that Lagrange  multipliers  could be represented by the responses of different types of olfactory neurons that solve the dual rather than primal representation problem.  and  0 i  Another motivation for relating  to neural activity can be derived from the Karush- Kuhn-Tucker theorem (KKTT) [9]. According to KKTT, at the allowed maximum of the dual Lagrangian (12), the Lagrangian contributions in equation (10) vanish, i.e.  and  5 u  i i ( )   0 , jx  0 j j (14) The former equality means that if A x   ij j 0 (inactive constraint), the corresponding 0 i  . KKTT therefore has significant implications from the standpoint of Lagrange multiplier i are identically equal zero (for inactive constraints) neural representations. That some values of makes the vector of responses sparse, which is one of the broadly observed features of neural responses [10; 11; 12] including in the olfactory system [13; 14; 15; 16]. The mapping between the dual problem and neural responses, if established, could explain the sparsity of neural responses as a corollary of KKTT. 0 i  and  H   , ) 2.6 Dual network. We will now describe the neural networks that solve the dual problem (12) and   (13). We associate vectors  and  with firing rates of two groups of neurons (cell types). The 0 j  are then satisfied automatically as firing rates cannot be negative. conditions We then assume that the neurons are connected into a network and that the Lyapunov function of  H     ) the network [cf. equation (12)]. To generate network equations from the Lyapunov function, for each neuron in the network, we define internal variables that can be viewed as total synaptic input current, for jb respectively. Consider example. For the following equations for a i j units, these currents will be denoted is proportional to the negative dual Lagrangian: u   i ia and i and (15)   A ij  a i u i       ( ( ) ( , , j  j ia and jb  W    ik k  b  b k j j   u A i ij j    i i u A ij i  1 (16) j k (17) (18) external excitatory drive equal to  i Here and throughout this paper we use with firing rates These cells are also connected to W u G u    ik i ik ik b a [ ] , ] [    j i   dx dt x / . The first equation describes inputs into cells ikW . Connectivity between cells is symmetric. i connected by weights . cells also receive j cells that are j cells with synaptic weights    . Equation (16) describes connected symmetrically to cells. For both cell types, their firing rates ( jb ) by rectifying threshold-linear relationships (18) ([ ]x connected to their inputs ( 0 0 for rewritten as forms of gradient descend  , j ) are x   x  ). It is straightforward to show that these equations can be   i is indeed the Lyapunov function of equations (15)-(18), i.e. a function To show that that is not increasing when our network evolves according to these equations, we observe that H    i H    j x  and [ ] x   for i and ia and iju A  ,   (19) A ij ) / u i ) / u i  b 0   , ( ( j i j ) / dt       i H  ) /  ( , i           j H  ) /  ( , j (  a  H   , )   (   , dH j  A x ij  j    i  a  i i   j  i   b  j j    i  f a a ) '( i 2 i  6 j   j  f b b '( ) j 2 j  0  (20) 0 f x ( ) x  [ ] , f 0   x  . Because '( ) Here , our network will minimize the Lyapunov function, and, by extension, maximize the dual Lagrangian (12). Because of the constraints imposed on the firing rates (18), these variables will stay non-negative in the course of this optimization, thus automatically satisfying the constraints imposed on the variables of the dual Lagrangian. We conclude therefore that our network, which includes only two cell types, can solve the dual constraint optimization problem, leading to the accurate computation of the molecular composition of a mixture in dual space. iu is to identify the set of primary variables The purpose of variables components of vector y that are larger than all others. To compute these variables one could use the network with winner-takes-all architecture, except one has to ensure that there are p winners. This network can be implemented by connecting N neurons with an unstructured global inhibition, in the manner described previously by us [17]: iy , i.e. the set of p ,    c  v i v i  u k y i k  , (   ) / dt dH  sgn( u i ensure that a given number of u - cells is active. v i ) . Here, c is the strength of global inhibition. By adjusting parameter c , one can Figure 1. Simulations with the network described. (A) The structure of the network. We propose that  cells implementing the dual representation of the concentration vector reside in the piriform cortex (PC). This is because PC is known to have extensive recurrent connectivity.  cells, which implement the non-negativity constraints on the concentration vector, could be related to the granule cells of the olfactory bulb [14; 18]. (B) Firing rate simulation of the dual model (blue) successfully identifies the non-zero components of the stimulus (top, red circles). Small false positive components of x result from our use of the elastic net measure and disappear as 0 . They can be removed by thresholding the reconstruction. In the simulation, we used p  60 . 3 Discussion Herein we propose a novel model for the intensity-invariant encoding of olfactory stimuli. According to the primacy model, an odor can be identified on the basis of the identities of the p strongest responding olfactory receptors. In the case of the human olfactory system, we have estimated p to be ~30. Because the primacy model relies on the relative rather than absolute strengths of receptor responses, the recovered stimulus is independent wrt the absolute stimulus 7 concentration and depends only on the relative concentrations of individual molecules in the mixture. Although we demonstrated this idea for the particular example of primacy coding, we suggest that this rule of neural relativity could be used to produce intensity invariant signal recovery for more complex conditions and in other modalities. We then attempted to formulate the solution of the decoding problem from the first principles. To implement primacy decoding, we cast this problem as a dual Lagrange problem, which required introducing two sets of Lagrange multipliers. The first set () enforced the primacy conditions, while the second set () ensured the non-negativity of the individual molecular concentrations. By assuming that the dual Lagrangian corresponds to a Lyapunov function of a network, we were able to derive the network structure. Interestingly, we found the cells were found to be recurrently connected, while  cells are connected to cells via inhibitory circuitry. These features are reminiscent of the cells in real olfactory networks. Pyramidal cells in the piriform cortex (PC) form extensive recurrent connections [19], while granule cells of the olfactory bulb receive inhibitory connections from PC [18]. The connection from the granule cells to the PC is indirectly inhibitory, because granule cells inhibit the mitral cells within the olfactory bulb through dendrodendritic synapses. Mitral cells are excitatory cells projecting to PC. These considerations allow us to place cells and related u cells into the piriform cortex and associate  cells with the granule cells of the olfactory bulb. Interestingly, the number of  cells matches the number of monomolecular components that the system can resolve, because these cells implement the conditions of non-negativity of the molecular concentrations. The number of granule cells found in the olfactory bulb is a few million, which matches with our estimate of the number of volatile molecules [18]. This coincidence provides a further argument in favor of the association of  cells with granule cells. Lengyel et al. [20] have proposed that the maximum a posteriori (MAP) solution to the inference problem of recovering a sparse N-dimensional odor vector x can be achieved in a low-dimensional measurement space, reflecting the known biology of olfactory processing. In their study, the compressed sensing problem was formulated as an 1l norm minimization of the odor vector x subject to the linear equality encoding constraint y = Ax and solved by considering the problem in dual space. The resulting generalized energy function for the network reflects the equality constraints and, as such, their proposed network implementation differs significantly from our solution, which relies on Karush-Kuhn-Tucker-type inequality conditions implementing primacy. Two features of dual networks are worth highlighting. First, according to Karush-Kuhn-Tucker theory, dual Lagrangians are optimized under constraints of non-negativity of the Lagrange coefficients. Neuronal responses can naturally enforce these constraints, because they are described by firing rates that cannot fall below zero. Secondly, due to KKTT [9], a large number of the Lagrange coefficients are bound to be zero, drawing an interesting parallel with SVMs. This observation is consistent with the observed sparsity of neuronal responses, both in olfaction [13; 14; 15; 16] and beyond [10; 11; 12; 21], further strengthening the possible association between neuronal responses and Lagrange coefficients. We therefore provide arguments in favor of relating neuronal responses and Lagrange multipliers implementing several constraints present in the stimulus. 8 4 References [1] S. Firestein, How the olfactory system makes sense of scents. Nature 413 (2001) 211- 8. [2] A. Koulakov, A. Gelperin, and D. Rinberg, Olfactory coding with all-or-nothing glomeruli. J Neurophysiol 98 (2007) 3134-42. [3] X. Zhang, and S. Firestein, The olfactory receptor gene superfamily of the mouse. Nat Neurosci 5 (2002) 124-33. [4] S. Kim, P.A. Thiessen, E.E. Bolton, J. Chen, G. Fu, A. Gindulyte, L. Han, J. He, S. He, B.A. Shoemaker, J. Wang, B. Yu, J. Zhang, and S.H. Bryant, PubChem Substance and Compound databases. Nucleic Acids Res 44 (2016) D1202-13. [5] A. Jinks, and D.G. Laing, A limit in the processing of components in odour mixtures. Perception 28 (1999) 395-404. [6] R.G. Baraniuk, Compressive sensing. Ieee Signal Proc Mag 24 (2007) 118-+. [7] D.L. Donoho, and J. Tanner, Sparse nonnegative solution of underdetermined linear equations by linear programming. P Natl Acad Sci USA 102 (2005) 9446-9451. [8] D.L. Donoho, and J. Tanner, Thresholds for the recovery of sparse solutions via L1 minimization. 2006 40th Annual Conference on Information Sciences and Systems, Vols 1-4 (2006) 202-206. [9] S.P. Boyd, and L. Vandenberghe, Convex optimization, Cambridge University Press, Cambridge, UK ; New York, 2004. [10] M.R. DeWeese, and A.M. Zador, Non-Gaussian membrane potential dynamics imply sparse, synchronous activity in auditory cortex. J Neurosci 26 (2006) 12206-18. [11] T. Hromadka, M.R. Deweese, and A.M. Zador, Sparse representation of sounds in the unanesthetized auditory cortex. PLoS biology 6 (2008) e16. [12] W.E. Vinje, and J.L. Gallant, Sparse coding and decorrelation in primary visual cortex during natural vision. Science 287 (2000) 1273-6. [13] L.M. Kay, and G. Laurent, Odor- and context-dependent modulation of mitral cell activity in behaving rats. Nat Neurosci 2 (1999) 1003-9. [14] A.A. Koulakov, and D. Rinberg, Sparse incomplete representations: a potential role of olfactory granule cells. Neuron 72 (2011) 124-36. [15] D. Rinberg, A. Koulakov, and A. Gelperin, Sparse odor coding in awake behaving mice. J Neurosci 26 (2006) 8857-65. [16] D.D. Stettler, and R. Axel, Representations of odor in the piriform cortex. Neuron 63 (2009) 854-64. [17] H. Sanders, B.E. Kolterman, R. Shusterman, D. Rinberg, A. Koulakov, and J. Lisman, A network that performs brute-force conversion of a temporal sequence to a spatial pattern: relevance to odor recognition. Front Comput Neurosci 8 (2014) 108. [18] G.M. Shepherd, The synaptic organization of the brain, Oxford University Press, Oxford ; New York, 2004. [19] K.M. Franks, M.J. Russo, D.L. Sosulski, A.A. Mulligan, S.A. Siegelbaum, and R. Axel, Recurrent circuitry dynamically shapes the activation of piriform cortex. Neuron 72 (2011) 49-56. [20] S.a.L. Tootoonian, M. , A Dual Algorithm for Olfactory Computation in the Locust Brain. Advances in Neural Information Processing Systems 27 (NIPS 2014) (2014). 9 [21] S.R. Lehky, T.J. Sejnowski, and R. Desimone, Selectivity and sparseness in the responses of striate complex cells. Vision Res 45 (2005) 57-73. 10
1611.10252
1
1611
2016-11-29T18:11:00
SeDMiD for Confusion Detection: Uncovering Mind State from Time Series Brain Wave Data
[ "q-bio.NC", "cs.AI", "cs.LG" ]
Understanding how brain functions has been an intriguing topic for years. With the recent progress on collecting massive data and developing advanced technology, people have become interested in addressing the challenge of decoding brain wave data into meaningful mind states, with many machine learning models and algorithms being revisited and developed, especially the ones that handle time series data because of the nature of brain waves. However, many of these time series models, like HMM with hidden state in discrete space or State Space Model with hidden state in continuous space, only work with one source of data and cannot handle different sources of information simultaneously. In this paper, we propose an extension of State Space Model to work with different sources of information together with its learning and inference algorithms. We apply this model to decode the mind state of students during lectures based on their brain waves and reach a significant better results compared to traditional methods.
q-bio.NC
q-bio
JMLR: Workshop and Conference Proceedings 1 (2016) 1–11 NIPS 2016, Time Series Workshop SeDMiD for Confusion Detection: Uncovering Mind State from Time Series Brain Wave Data Jingkang Yang International School Beijing University of Posts and Telecommunications Beijing, 100876, China Haohan Wang Language Technologies Institute School of Computer Science Carnegie Mellon University Pittsburgh, PA, 15206, the USA [email protected] [email protected] Jun Zhu State Key Laboratory of Intelligent Technology and Systems Tsinghua National Laboratory for Information Science and Technology Tsinghua University Beijing, 100084, China [email protected] Eric P. Xing Machine Learning Department School of Computer Science Carnegie Mellon University Pittsburgh, PA, 15206, the USA Editor: List of editors' names [email protected] Abstract Understanding how brain functions has been an intriguing topic for years. With the recent progress on collecting massive data and developing advanced technology, people have be- come interested in addressing the challenge of decoding brain wave data into meaningful mind states, with many machine learning models and algorithms being revisited and devel- oped, especially the ones that handle time series data because of the nature of brain waves. However, many of these time series models, like HMM with hidden state in discrete space or State Space Model with hidden state in continuous space, only work with one source of data and cannot handle different sources of information simultaneously. In this paper, we propose an extension of State Space Model to work with different sources of information together with its learning and inference algorithms. We apply this model to decode the mind state of students during lectures based on their brain waves and reach a significant better results compared to traditional methods. Keywords: Sequence Data based Mind-Detecting (SeDMiD) Model, Time Series, Brain Wave, Mind State Reading c(cid:13) 2016 J. Yang, H. Wang, J. Zhu & E.P. Xing. Yang Wang Zhu Xing 1. Introduction Understanding how human brain functions has been an attractive research question in recent years Mitchell et al. (2008). One important progress is on collecting a large amount of brain wave data with different technologies, like fMRI Wehbe et al. (2014), MEG Sudre et al. (2012) and EEG Wang et al. (2013). The nature of these data collecting technologies have introduced a variety of substantial challenges in understanding these data with machine learning techniques. For example, MEG and EEG technologies can describe the brain with considerable temporal granularity, but with a relatively low spatial resolution. Therefore, machine learning techniques that can handle temporal dependencies are highly appreciated. Fortunately, in recent years, there is an increasing trend towards the use models to work with time series problem. For example, Khaleghi and Ryabko (2013) find the points in time where the probability distribution generating the data has changed given a heterogeneous time-series sample. Anava et al. (2013) use regret minimization techniques to develop ef- fective online learning algorithms for predicting a time series using autoregressive moving average model. Alon et al. (2003) fit a finite mixture of HMMs in motion data, using the expectation maximization (EM) framework, aiming at discovering groupings of similar object motions that were observed in a video collection. Also, Hidden Markov Models or other graph-based methods are often used for the purpose like speech recognition, pattern recognition and neural networks, similar to Nahar et al. (2016), Schwenk (1999), Bar-Joseph (2004) and LeCun and Bengio (1995). In a more rigorous setting, a lot more theoretical questions are being asked and solved for time series machine learning these days. For exam- ple, Khaleghi and Ryabko (2013, 2014) derive theories for estimation of highly dependent time series data. Kuznetsov and Mohri (2014, 2015) push the theoretical work further for non-stationary time series problems. Rakhlin and Sridharan (2013); ? naturally combines the problem of analyzing time series data with online learning, which opens the door to a whole area of new problems. With the guidance of previous work on time series data. In this paper, we present Sequence Data based Mind-Detecting (SeDMiD) Model , a novel time series method to un- cover the state of brain using more than a typical source of EEG/MEG recordxings. Others sources like video recordings or audio data can be supplemented for better performance on brain state estimation. We also develop the learning algorithm for SeDMiD based on sparsity regularized linear system, and the inference algorithm as an extension of Viterbi algorithm under Gaussian assumption for continuous space. The results show that, the performance of SeDMiD can uncover the mind state based on brain waves with a significant better results than traditional methods. Our contribution of this paper is three-fold: • We propose a SeDMiD that can analyze time series brain wave data with extra sources of information. • We improve the existing vertibi algorithm to enable the inference of SeDMiD model. • We show the possibility of deciphering students' mind state of understanding lectures with brain wave data. 2 SeDMiD for Confusion Detection The rest of this paper is organized as following. Section 2 describes some work that others did to solve the prediction problem using MEG data. Section 3 raises the novel model and describe its learning and inference method, while section 4 shows the result of our experiment. Finally, Section 5 concludes and suggests future work. 2. Related Work There are many works implementing time series technology to deal with the problem about MEG/EEG data. It is worth noting that both electroencephalography (EEG) and mag- netoencephalography (MEG) provide a more direct measure of the electrical activity in the brain, as is described professionally in the work of Michel (2009) and Proudfoot et al. (2014). He et al. (2008) measure the difference in electric potentials on the scalp and cap- tures high frequency oscillations on the millisecond timescale that is most relevant for the characterisation of cognitive processes. There are many existing work to do experiments and discover new knowledge in several fields. Moghadamfalahi et al. (2015) use abstract- noninvasive EEG-based braincomputer interfaces (BCI) for intent detection, specifically for EEG-based BCI typing systems. Phillips et al. (1997) have developed a Bayesian framework for image estimation from combined MEG/EEG data. EEG signal is a kind of voltage signal that can be measured on the surface of the scalp, arising from large areas of coordinated neural activity manifested as synchronization (groups of neurons firing at the same rate), described by Niedermeyer and da Silva (2005). This neural activity varies as a function of development, mental state, and cognitive activity, and the EEG signal can measurably detect such variation Marosi et al. (2002), Lutsyuk et al. (2006), Berka et al. (2007), Wang et al. (2013) which in turn are important for and predictive of learning Baker et al. (2010). On the other hand, hidden Markov model is also widely implemented to make best use of MEG/EEG recordings. Rukat et al. (2016) analyses the temporal and spatial dynamics of physiological substrate of cognitive processes as measured by EEG, with a hidden Markov model. Liu et al. (2010) combine kernel principal component analysis (KPCA) and HMM to differentiate mental fatigue states with the help of EEG data. Ko and Sim (2011) describe a procedure of classification of motor imagery EEG signals using HMM, which can tell the person is performing left, right hand or foots motor imagery based on the current EEG recordings. What is more, another bio-signal named electrocardiogram (ECG) is also suitable for Markov model. Coast et al. (1990) and Andreao et al. (2006) describe a new approach to ECG arrhythmia analysis based on HMM. Inspired by these work, we propose a novel model to analyze time series brain wave data with extra sources of information. 3. SeDMiD Model and Its Learning and Inference Algorithm We propose a novel state-space model for inferring people's state using several sources of information simultaneously including MEG/EEG recordings, named Sequence Data based Mind-Detecting (SeDMiD) Model. In this section, we will first introduce the model, and then show its learning and inference algorithm respectively. 3 Yang Wang Zhu Xing Figure 1: The proposed model. For the time-stamp at i, Mi corresponds to EEG/MEG recording and Ci corresponds to mind states (e.g. happy/sad, confused/non- confused). Si is extra source information (e.g. video/audio data). A, B, E, F and G are stationary transition matrices that describe the dynamics among states in the graph. 3.1. SeDMiD The aim of our work is to raise a model for brain-state estimation using MEG/EEG record- ings while considering other essential time-series sources at the same time. A pictorial representation of our model can be found in Figure 1. Notations of this paper is illustrated in the caption of Figure 1. In the SeDMiD model, we firstly assume that the extra time-series sources can match exactly with brain wave signals in aspect of time stamps, which means Si, Ci and Mi are sampled at the same time. And we simplify the mind state inferring task as a binary decision of states Ci, indicating that people is either happy or sad, or students are either confused or not. We also assume that MEG/EEG recordings Mi have linear dependence on current external source and current mind state, since brain signal is easily affected by people's internal mind situation and outside influence on them Baker et al. (2010). Additionally, current mind state Mi depends linearly on previous status and current video contents, because brain state often has a close connection with its state few seconds ago, and other sources like videos also affect the mind state. Finally, the complementary source is a continuous time-series process, thus current recording is dependent on the former one, and to simplify the model, we assume linear dependency again. Further, we assume that the supplemental source data {Si}n i=1 can be described by Gaussian distribution. We also assume that mind states {Ci}n i=1 is in form of Gaussian distribution. S1 ∼ N (µs, σs) Si ∼ N (A · Si−1, Λi) C1 ∼ N (µc, σc) 4 SeDMiD for Confusion Detection Ci ∼ Nl(E · Si + B · Ci−1), Ωi  Considering the assumption that only on-and-off state exist in space C, we use the function l() to map the score of positive state (E · Si + B · Ci−1) into two dimensions above. (cid:20) 1 (cid:21) 1+ex ex 1+ex l(x) = Thus brain wave recordings {Mi}n linear combination of {Si}n co-variance is Σi. i=1 and {Ci}n i=1 also follows Gaussian distribution because it is the i=1, as the mean of Mi is G · Si + F · Ci and Mi ∼ N (G · Si + F · Ci, Σi) The practical meaning of notations in the model is described by Table 1. Table 1: Practical Meaning of Notations in SeDMid model Notation Description Si Ci Mi A B E F G Features of supplemental source information at time i Mind States at time i Brain wave signal recordings (EEG/MEG) at time i Linear Relationship between adjacent supplemental source information Linear Relationship between adjacent mind states Linear Relationship between supplemental source and mind state Linear Relationship between mind state and brain wave signal Linear Relationship between supplemental source and brain wave signal 3.2. Learning Here we introduce the parameter learning algorithm of A, B, E, F and G for SeDMiD. Because of the nature of the task, a supervised training training procedure, with S, C, M known, is sufficient. At training, at time-stamp i, we can observe MEG/EEG recording Mi, supplemental source data Si and current mind state Ci. Thus, parameter learning is a maximum likelihood estimation (MLE) problem, in which case linear regression is the solution. However, supplemental source data generally comes with a higher dimension than response variable space. Therefore, sparsity regularize is required for transition matrix S by Equation (3) in a sparse form. Mi =(cid:2)G F(cid:3)(cid:20)Si Ci =(cid:2)B E(cid:3)(cid:20) Ci Ci (cid:21) (cid:21) Si+1 Si+1 = A · Si 5 (1) (2) (3) Yang Wang Zhu Xing 3.3. Inference In inference, observation only contains brain wave recordings, so we want to estimate mind state Ci and even further inference complementary data Si. With the observation Mi and five transition matrices that learned in the learning phase, we formulate an inference method to find the best sequence of mind states C1, C2, . . . , CT and features of complementary data S1, S2, . . . , ST . A natural choice for inference on the model is Viterbi algorithm Forney (1973). However, the problem is that the state space S and M are infinite, and we assume that Si and Mi follow Gaussian distribution, we formulate the inference mathematically in Gaussian form. Consider the Viterbi function at the time-stamp t is calculated by Equation (4), given by the factorization of graphic model shown in Figure 3.1. Vt(St, Ct) = max St−1,Ct−1 [P (MtSt, Ct)P (StSt−1)P (CtSt, Ct−1)Vt−1(St−1, Ct−1)] (4) To begin with, we consider to calculate the Viterbi function V1 at the start point of the whole process when t = 1. V1(S1, C1) = P (M1S1, C1)P (S1)P (C1S1) = N (G · S1 + F · C1, Σ) · N (µs, σs) · N (l(E · µs), σc) (5) According to the property of multiplication in Gaussian distribution, we obtain that V1 is also in form of Gaussian with its mean and variance shown in (6) and (7) respectively. (cid:20)σ−1 (cid:20) s 0 (cid:21) (cid:21) (cid:20)GT (cid:21) F T Σ−1(cid:2)G F(cid:3)−1  (cid:21) (cid:20)GT Σ−1M1 F T + c + 0 σ−1 σ−1 s µs c l(E · µs) σ−1 ΣV1 = µV1 = ΣV1 (6) (7) For now, the mean and covariance of Viterbi function for the beginning point have been calculated, and in order to know all the Viterbi function at every time-stamp, we find when t (cid:62) 2: Vt(St, Ct) = max St−1,Ct−1 P (CtSt, Ct−1)P (StSt−1)Vt−1(St−1, Ct−1)P (MtSt, Ct) (8) Also, we can calculate the mean and covariance since for time t, St and Ct are in the form of Gaussian distribution. ΣV2 = µV2 = ΣV2 F T Φ−1 + (cid:20)GT (cid:21) Φ−1 · µV1 + (cid:20)(A−1)T Σ−1(cid:2)G F(cid:3)−1  (cid:20)GT (cid:21) (cid:20)A−1 (cid:21) Σ−1M2 F T Σ−1 (B−1)T V1 0 0 0 0 B−1 (9) (10) (11) (cid:21)−1 where (cid:20)Λ−1 0 0 Ω−1 (cid:21) Φ = + 6 SeDMiD for Confusion Detection Once the means and co-variances are calculated for all supplemental source data and mind state sequence, the best sequence of S and C can be inferred by calculating backwards. So we start the inference at the end of the time sequence. (cid:21) (cid:20)ST CT = arg max ST ,CT N (µVT , ΣVT ) (12) Note that the mean of the final state is calculated exactly, we proceed to infer the former state that lead to the current one, so the formula to calculate the former state based on current state is shown in (13).(cid:20)ST−1 (cid:21) (cid:20)(E · A)T (cid:21) (cid:20)(E · A)T Σ−1 Σ−1 µVT−1 + CT−1 where γ = τ VT−1 VT−1 τ = BT + BT = arg max N (γ, τ ) ST −1,CT −1 Ω−1(cid:2)E · A B(cid:3) + (cid:21) (cid:21) (cid:20)ST Ω−1 + (cid:21)−1 (cid:20)AT Λ−1A 0 (cid:21)−1 (cid:21)(cid:20)ST (cid:20)AT Λ−1A 0 0 0 0 0 CT CT (13) (14) (15) Finally, by calculate all the Si and Ci from end to the first one, we can infer mind states for each time-stamp t ∈ {1, . . . , T}. 4. Experiment To solve the problem that we raised in Section 4.1, we need to extract features of lecture videos. Using SeDMid model that we raise in Section 3, 4.1. Experiment Setting In our experiment, we set up a task to estimate students' mind states in a given lecture session, finding out they are confused or not with the data from Wang et al. (2013). Every participants is asked to watch 10 lecture videos, the length of which is around 2 minutes.The two-minute period is called an 'experiment period'. In all there are 10 students participate in the experiments so there are 100 data points in total. During every experiment period, their EEG signals are recorded in the frequency of 1 Hz, and EEG signals here have 11 features, including Proprietary measure of mental focus, 1-3 Hz of power spectrum and so on. Students are asked to annotate whether they are confused or not based on every whole experiment period, and every second in the period by the annotation is noted. The beginning and the end of every experiment is cut off with consideration of reducing noise, only left the middle 112 seconds for analysis. Finally, lecture videos are served as supplemental source data. 4.2. Video Feature Extraction We use the tool kit OpenCV developed by Bradski and Kaehler (2008) to extract video features like optical features and object movement information, and openSMILE developed 7 Yang Wang Zhu Xing by Eyben et al. (2010) to extract features in audio data, such as lecturer's speech speed and intonation. As a result, we obtain video image features with 1440 dimensions while audio features have 6669 dimensions, thus we get 8109 dimensions for video features in total. Since both MEG recordings and students' status vector is collected in the frequency of 1 Hz, we sample the feature vector every 1 second in order to alignment the data for SeDMiD model. 4.3. Performance Comparison We use the simple state space model (SSM) as baseline, which only make use of EEG data, and current mind state only depends on the former state, while current EEG recording depends on current mind state. SSM does not make use of video features. We also compare our model with logistic regression, which employ brain wave signals without other sources. As is shown in the ROC curve in Figure 2, we find that our SedMid model outperform the simple HMM model and logistic regression, with the accuracy of ours reaches 87.76% while simple HMM only gets 53% and logistic regression 60%. In the experiment, We find that errors that SeDMiD makes always exist at the beginning of experiment period, and it will lead to correct result in few time. It is intuitive that SeDMiD model often makes mistakes at beginning of every experiment period because it can perform better with more data gets concluded. Figure 2: The ROC curve for comparing the performance of SeDMiD model, Simple State Model (SSM) and logistic regression. SeDMiD greatly outperforms the others. 8 SeDMiD for Confusion Detection 5. Conclusion In this work, we propose an novel state space model called Sequence Data based Mind- Detecting (SeDMiD) Model, which analyzes time series brain wave data with extra sources of information, improving the existing vertibi algorithm to enable the inference. We evaluate the effectiveness of SeDMiD model by comparing with simple Markov model and logistic regression. The performance of our model has a 30% higher in accuracy than the normal one. Apart from proposing the SeDMiD model, our contribution includes showing the possi- bility of deciphering students' mind state of understanding lectures with brain wave data. This work is also useful in real world implementation, since teachers can modify their teach- ing strategy based on audiences' status if it can be inferred. Acknowledgments This work was supported by International School, Beijing University of Posts and Telecom- munications. We thank Hao Ding for advising the method of extracting videos' features and dormitory administrator in Student Building 5, BUPT, for providing electronic power for running experiment at night. References Jonathan Alon, Stan Sclaroff, George Kollios, and Vladimir Pavlovic. Discovering clusters in motion time-series data. In Computer Vision and Pattern Recognition, 2003. Proceedings. 2003 IEEE Computer Society Conference on, volume 1, pages I–375. IEEE, 2003. Oren Anava, Elad Hazan, Shie Mannor, and Ohad Shamir. Online learning for time series prediction. In COLT, pages 172–184, 2013. Rodrigo Varejao Andreao, Bernadette Dorizzi, and J´erome Boudy. Ecg signal analysis through hidden markov models. IEEE Transactions on Biomedical engineering, 53(8): 1541–1549, 2006. Ryan SJd Baker, Sidney K D'Mello, Ma Mercedes T Rodrigo, and Arthur C Graesser. Better to be frustrated than bored: The incidence, persistence, and impact of learners cognitive–affective states during interactions with three different computer-based learning environments. International Journal of Human-Computer Studies, 68(4):223–241, 2010. Ziv Bar-Joseph. Analyzing time series gene expression data. Bioinformatics, 20(16):2493– 2503, 2004. Chris Berka, Daniel J Levendowski, Michelle N Lumicao, Alan Yau, Gene Davis, Vladimir T Zivkovic, Richard E Olmstead, Patrice D Tremoulet, and Patrick L Craven. Eeg corre- lates of task engagement and mental workload in vigilance, learning, and memory tasks. Aviation, space, and environmental medicine, 78(Supplement 1):B231–B244, 2007. Gary Bradski and Adrian Kaehler. Learning OpenCV: Computer vision with the OpenCV library. " O'Reilly Media, Inc.", 2008. 9 Yang Wang Zhu Xing Douglas A Coast, Richard M Stern, Gerald G Cano, and Stanley A Briller. An approach IEEE Transactions on to cardiac arrhythmia analysis using hidden markov models. biomedical Engineering, 37(9):826–836, 1990. Florian Eyben, Martin Wollmer, and Bjorn Schuller. Opensmile: the munich versatile and fast open-source audio feature extractor. In Proceedings of the 18th ACM international conference on Multimedia, pages 1459–1462. ACM, 2010. G David Forney. The viterbi algorithm. Proceedings of the IEEE, 61(3):268–278, 1973. Biyu J He, Abraham Z Snyder, John M Zempel, Matthew D Smyth, and Marcus E Raichle. Electrophysiological correlates of the brain's intrinsic large-scale functional architecture. Proceedings of the National Academy of Sciences, 105(41):16039–16044, 2008. Azadeh Khaleghi and Daniil Ryabko. Nonparametric multiple change point estimation in highly dependent time series. In International Conference on Algorithmic Learning Theory, pages 382–396. Springer, 2013. Azadeh Khaleghi and Daniil Ryabko. Asymptotically consistent estimation of the number of change points in highly dependent time series. In ICML, pages 539–547, 2014. Kwang-Eun Ko and Kwee-Bo Sim. Hsa-based hmm optimization method for analyzing eeg pattern of motor imagery. Journal of Institute of Control, Robotics and Systems, 17(8): 747–752, 2011. Vitaly Kuznetsov and Mehryar Mohri. Generalization bounds for time series prediction with non-stationary processes. In International Conference on Algorithmic Learning Theory, pages 260–274. Springer, 2014. Vitaly Kuznetsov and Mehryar Mohri. Learning theory and algorithms for forecasting non- In Advances in Neural Information Processing Systems, pages stationary time series. 541–549, 2015. Yann LeCun and Yoshua Bengio. Convolutional networks for images, speech, and time series. The handbook of brain theory and neural networks, 3361(10):1995, 1995. Jianping Liu, Chong Zhang, and Chongxun Zheng. Eeg-based estimation of mental fa- tigue by using kpca–hmm and complexity parameters. Biomedical Signal Processing and Control, 5(2):124–130, 2010. NV Lutsyuk, EV ´Eismont, and VB Pavlenko. Correlation of the characteristics of eeg potentials with the indices of attention in 12-to 13-year-old children. Neurophysiology, 38 (3):209–216, 2006. Erzs´ebet Marosi, Oscar Baz´an, Guillermina Yanez, Jorge Bernal, Thal´ıa Fern´andez, Mario Rodr´ıguez, Juan Silva, and Alfonso Reyes. Narrow-band spectral measurements of eeg during emotional tasks. International Journal of Neuroscience, 112(7):871–891, 2002. Christoph M Michel. Electrical neuroimaging. Cambridge University Press, 2009. 10 SeDMiD for Confusion Detection Tom M Mitchell, Svetlana V Shinkareva, Andrew Carlson, Kai-Min Chang, Vicente L Malave, Robert A Mason, and Marcel Adam Just. Predicting human brain activity associated with the meanings of nouns. science, 320(5880):1191–1195, 2008. Mohammad Moghadamfalahi, Umut Orhan, Murat Akcakaya, Hooman Nezamfar, Melanie Fried-Oken, and Deniz Erdogmus. Language-model assisted brain computer interface for typing: a comparison of matrix and rapid serial visual presentation. IEEE Transactions on Neural Systems and Rehabilitation Engineering, 23(5):910–920, 2015. Khalid MO Nahar, Mohammed Abu Shquier, Wasfi G Al-Khatib, Husni Al-Muhtaseb, and Moustafa Elshafei. Arabic phonemes recognition using hybrid lvq/hmm model for continuous speech recognition. International Journal of Speech Technology, pages 1–14, 2016. Ernst Niedermeyer and FH Lopes da Silva. Electroencephalography: basic principles, clinical applications, and related fields. Lippincott Williams & Wilkins, 2005. James W Phillips, Richard M Leahy, John C Mosher, and Bijan Timsari. Imaging neural activity using meg and eeg. IEEE Engineering in Medicine and Biology Magazine, 16(3): 34–42, 1997. Malcolm Proudfoot, Mark W Woolrich, Anna C Nobre, and Martin R Turner. Magnetoen- cephalography. Practical neurology, pages practneurol–2013, 2014. Alexander Rakhlin and Karthik Sridharan. Online learning with predictable sequences. In COLT, pages 993–1019, 2013. Tammo Rukat, Adam Baker, Andrew Quinn, and Mark Woolrich. Resting state brain arXiv preprint networks from eeg: Hidden markov states vs. classical microstates. arXiv:1606.02344, 2016. Holger Schwenk. Using boosting to improve a hybrid hmm/neural network speech rec- In Acoustics, Speech, and Signal Processing, 1999. Proceedings., 1999 IEEE ognizer. International Conference on, volume 2, pages 1009–1012. IEEE, 1999. Gustavo Sudre, Dean Pomerleau, Mark Palatucci, Leila Wehbe, Alona Fyshe, Riitta Salmelin, and Tom Mitchell. Tracking neural coding of perceptual and semantic fea- tures of concrete nouns. NeuroImage, 62(1):451–463, 2012. Haohan Wang, Yiwei Li, Xiaobo Hu, Yucong Yang, Zhu Meng, and Kai-min Chang. Using eeg to improve massive open online courses feedback interaction. In AIED Workshops, 2013. Leila Wehbe, Brian Murphy, Partha Talukdar, Alona Fyshe, Aaditya Ramdas, and Tom Mitchell. Simultaneously uncovering the patterns of brain regions involved in different story reading subprocesses. PloS one, 9(11):e112575, 2014. 11
1612.02243
1
1612
2016-12-07T13:47:58
How structure sculpts function: unveiling the contribution of anatomical connectivity to the brain's spontaneous correlation structure
[ "q-bio.NC" ]
Intrinsic brain activity is characterized by highly structured co-activations between different regions, whose origin is still under debate. In this paper, we address the question whether it is possible to unveil how the underlying anatomical connectivity shape the brain's spontaneous correlation structure. We start from the assumption that in order for two nodes to exhibit large covariation, they must be exposed to similar input patterns from the entire network. We then acknowledge that information rarely spreads only along an unique route, but rather travels along all possible paths. In real networks the strength of local perturbations tends to decay as they propagate away from the sources, leading to a progressive attenuation of the original information content and, thus, of their influence. We use these notions to derive a novel analytical measure, $\mathcal{T}$ , which quantifies the similarity of the whole-network input patterns arriving at any two nodes only due to the underlying topology, in what is a generalization of the matching index. We show that this measure of topological similarity can indeed be used to predict the contribution of network topology to the expected correlation structure, thus unveiling the mechanism behind the tight but elusive relationship between structure and function in complex networks. Finally, we use this measure to investigate brain connectivity, showing that information about the topology defined by the complex fabric of brain axonal pathways specifies to a large extent the time-average functional connectivity observed at rest.
q-bio.NC
q-bio
How structure sculpts function: unveiling the contribution of anatomical connectivity to the brain's spontaneous correlation structure R. G. Bettinardi,1, 2, a) G. Deco,1, 2, 3 V. M. Karlaftis,4 T. J. Van Hartevelt,5, 6 H. M. Fernandes,5, 6 Z. Kourtzi,4 M. L. Kringelbach,5, 6 and G. Zamora-L´opez1, 2, b) 1)Center for Brain and Cognition, Universitat Pompeu Fabra, Barcelona, Spain. 2)Department of Information and Communication Technologies, Universitat Pompeu Fabra, Barcelona, Spain. 3)Instituci´o Catalana de la Recerca i Estudis Avan¸cats, Universitat Pompeu Fabra, Barcelona, Spain. 4)Department of Psychology, University of Cambridge, United Kingdom. 5)Department of Psychiatry, University of Oxford, United Kingdom 6)Center for Music in the Brain, Aarhus University, Aarhus, 8000 Aarhus C, Denmark Intrinsic brain activity is characterized by highly structured co-activations between different regions, whose origin is still under debate. In this paper, we address the question whether it is possible to unveil how the underlying anatomical connectivity shape the brain's spontaneous correlation structure. We start from the assumption that in order for two nodes to exhibit large covariation, they must be exposed to similar input patterns from the entire network. We then acknowledge that information rarely spreads only along an unique route, but rather travels along all possible paths. In real networks the strength of local perturbations tends to decay as they propagate away from the sources, leading to a progressive attenuation of the original information content and, thus, of their influence. We use these notions to derive a novel analytical measure, T , which quantifies the similarity of the whole-network input patterns arriving at any two nodes only due to the underlying topology, in what is a generalization of the matching index. We show that this measure of topological similarity can indeed be used to predict the contribution of network topology to the expected correlation structure, thus unveiling the mechanism behind the tight but elusive relationship between structure and function in complex networks. Finally, we use this measure to investigate brain connectivity, showing that information about the topology defined by the complex fabric of brain axonal pathways specifies to a large extent the time-average functional connectivity observed at rest. I. INTRODUCTION In the last two decades a large body of research has demonstrated that spontaneous brain activity forms structured patterns of consistent co-activations across different subsets of brain regions 1 -- 10. Contrary to what was somehow implicitly assumed, intrinsic brain activity cannot be considered as a simple sum of rather unpre- dictable and noisy fluctuations. Spontaneous collective dynamics of different brain regions, measured either with EEG, fMRI or MEG, can be clustered into highly orga- nized and reproducible spatial patterns, referred to as resting-state networks (RSNs), that strikingly resemble those activations observed during the performance of dif- ferent tasks7,11,12. Furthermore, growing evidences sug- gest that this spontaneous large-scale structure is char- acterized by a marked temporal organization, mirrored by recurrent alternations of subnetworks that concur in generating a rich dynamical repertoire at different time scales13,14. Although the origin and the purpose of spon- taneous brain dynamics are still under debate, there is wide agreement that mental states and brain malfunc- tion alter the patterns of spontaneous activity, e.g. the a)Electronic mail: [email protected] b)Electronic mail: [email protected] dynamical repertoire of the brain at rest tends to decrease during sleep15 and under anaesthesia16 -- 18. From its discovery, great efforts have been invested in trying to reproduce resting-state brain activity through the use of computational models, in order to obtain a mechanistic explanation of this intriguing but elusive phenomenon. Early models based on the structural con- nectomes of cats and macaques extensively explored the emerging patterns of correlations in those networks at dif- ferent spatial and temporal scales19 -- 21. With the arrival of structural human connectomes obtained through trac- tography, computational models could finally attempt to fit the empirical correlation structure observed via resting-state fMRI22 -- 26. Despite these attempts, we still lack of a precise understanding of the relationship be- tween the shape of the brain's connectome and the emer- gent patterns of correlations observed during rest. One of the major reasons is that the collective dynamics of a network do not only depend on the shape of the un- derlying connectivity, but also on the model chosen to simulate the local dynamics of the cortical regions26,27. In this paper we aim at unveiling what is the precise contribution of the anatomical connectivity on the cor- relation structure observed during rest. To this aim, we first need to derive a theoretical estimate of the expected correlation between nodes due to the network's topology alone; therefore we will try to avoid, as much as possible, the contribution of other factors. To do so, in Section II we introduce a novel graph theoretical quantity measur- ing the topological similarity of the entire input profiles that two nodes receive from the whole network. This measure, that we named topological similarity, is a gener- alization of the concept of matching index that explicitly accounts for the fact that in networks information trav- els along all possible paths, not only along the shortest ones, and that information content tends to decay as it moves away from its source28 -- 31. This measure is based on the concept of network's communicability introduced by Estrada and Hatano 32, a function that quantifies the strength of the influence that one node exerts over an- other through all paths of any length assuming an ex- ponential decay of influence with path length. We then demonstrate that the topological similarity function, al- tough based on pure topological information about the underlying path structure of the network, can indeed be used to approximate the expected correlation structure of networks of time-varying coupled units. In Section III we systematically investigate the con- tribution of three topological primitives: the weight of the links, the length of the path and the presence of redundant alternative paths through which information can travel. Topological primitives are fundamental fea- tures underlying the network architecture which deter- mine how influence spreads, thus sculpting the similarity of the input profiles of nodes. To do so, we study simple graphs in which these primitive features can be manip- ulated. We show how these three features are in fact of primary importance in modulating the strength of the influence and the relative topologies into which different nodes are embedded. Finally, in Section IV we investigate the spontaneous correlation structure of the human brain, demonstrat- ing that taking into account the similarity of the whole- network influences shaped by the underlying anatomi- cal topology it is possible to understand and predict the large-scale functional organization observed during rest. II. HOW TOPOLOGY SCULPTS THE CORRELATION STRUCTURE OF NETWORKS In order to properly address the contribution of net- work's topology on the emerging spontaneous correlation structure, we first need to acknowledge that the collective behaviour of a set of coupled dynamical units depends on three principal ingredients: (i) the structure of the net- work, (ii) the local dynamics of the nodes and (iii) the coupling function determining how information is passed from one node to another. In fact, for a fixed network, changing the local dynamical model of the nodes and the coupling function usually leads to different collective dynamics26,27,33. Therefore, in order to estimate the con- tribution of structure alone we need to set apart the role of the other two factors. Typically, the activity of two nodes exhibits statistical 2 dependence either if they are connected by means of a direct link, or if the aggregate inputs they receive from the entire network are similar, independently of whether there is a link between the two or not. Because infor- mation in a network rarely travels exclusively along the shortest paths34,35 but instead diffuses along the whole network, we realise that the total influence of one node over another mainly depends on three topological fea- tures: (i) the strength of the coupling between them, usu- ally represented by the weights of the links, (ii) the graph distance between the two nodes and (iii) the propagation of the influence along multiple alternative paths36,37. We refer to these three topological features as the topologi- cal primitives because, in combination, they encode the shape of the network's path structure. As such the anal- ysis of the building blocks underlying larger complex net- works is a necessary step to understand the contribution of network structure to the emergent activity. In general, the influence of a direct link is greater than the influence exerted over longer paths, as the latter is mediated through third nodes. In fact, in real systems the "power" of the signals or their information content naturally decays along the path28 -- 31, unless there exists an active mechanism which amplifies the incoming signal at the cost of energy. Following the same rationale, it is unlikely that influence or information from one node to another propagates only along a single, selected path, unless there are specific gating mechanisms controlling the detailed routing of information over all existing paths. The total number of paths (of all lengths) between any two nodes in a network is in fact infinite. It is well known that the total number of paths ml ij of length l between nodes i and j in a graph grows with l. This number is given exactly by the lth power of the adjacency matrix A, mij = (Al)ij 38,39. Thus, the total number of paths leaving from node i and arriving at node j is given by the sum: ∞(cid:88) l=0 (Al)ij = 1 + Aij + (A2)ij + (A3)ij + . . . (1) which the series(cid:80)∞ This number typically diverges and thus, for the dy- namics within a network to remain bounded, the amount of influence needs to decay faster with the length than the growth in the number of paths. Mathematically, the problem consists in finding a set of coefficients {kl} for l=0 kl Al converges for any adjacency matrix, A. While the solution to this problem is not unique, Estrada and Hatano 32 proposed an exponential decay of the influence with path length and introduced the communicability measure C. The communicability function thus corresponds to the matrix exponential of A, which can be expanded into a series of powers with coefficients kl = 1/ l!: C ≡ eA = ∞(cid:88) l=0 Al l! = 1 + A + A2 2! + A3 3! + . . . (2) From a physical perspective, the communicability is analogous to the Green's function of the network32,40 and expresses how local perturbations propagate along the system. Perturbation of a given node's dynamics in fact first propagates to its direct neighbours, affecting their activity; the activity of the neighbours will in turn prop- agate to their neighbours (as well as back to the node that was perturbed in the first place), and interact with their intrinsic local dynamics. This simple propagation mechanism implies that the effect of a local perturbation could be perceived also by distant nodes but attenuated and modulated by the dynamics of each node along the route. Due to its correspondence to the Green's function, the communicability can be tuned using a constant global coupling parameter that uniformly scales the weights of all links in A40,41, allowing to search through the emerg- ing collective dynamics over multiple scales. The gener- alised, tunable, communicability is then: ∞(cid:88) l=0 C = egA = glAl l! = 1 + gA + g2A2 2! + g3A3 3! + . . . (3) When g is weak, perturbations quickly decay, produc- ing local correlations only around the node's neighbour- hood. As g grows perturbations propagate deeper into the network, giving raise to stronger correlations over more distant nodes. Zamora-L´opez et al. 41 have shown that considering the communicability as the propagator kernel for the diffusion of Gaussian noise along the net- work, it reveals an equivalent correlation structure as net- works of generic and widely used models, e.g. Kuramoto oscillators and neural masses (see Supplementary Infor- mation in Ref. 41). Under these assumptions (Gaussian white noise sources and exponential decay), in Ref 41 it was shown that the covariance matrix of the system, Σ, can be analytically estimated as Σ = C · CT where CT is the transposed communicability. The cross-correlation matrix R of the system is then calculated, as usual, nor- malising the rows of the covariances Σij by the auto- covariances Σii as, Rij = Σij/Σii. Despite the merit of being analytical, this estimate of the network's cross- correlation matrix still relied on assuming very simple dynamical model, the diffusion of Gaussian white noise, and did not fully disentangle the unique contribution of the topology. In the present work we want to address the question whether it is possible to estimate the most likely correlation structure that a network gives rise, based only on its topological properties. As mentioned before, it is legitimate to assume that the activities of two nodes will exhibit statistical dependence either if they are connected by means of a direct link, or if the aggregate input they receive from the entire network is similar. With this in 3 mind, we realised that the column vectors cj of the com- municability matrix C indeed represent the input profile of the influences node j receives from all nodes in the network along all possible paths, including the influence a node has on itself due to recurrent paths. Therefore, it should be possible to predict analytically the magni- tude of the expected correlation between any two nodes in a network by comparing the similarity of their input profiles. The resemblance between two input profiles can in principle be calculated using any measure of similarity between multidimensional vectors. Here we choose the cosine similarity because it returns results bounded be- tween −1 and 1, equivalent to the cross-correlation mea- sure we aim at comparing. Consequently, we define the topological similarity, Tij, between two nodes as the co- sine of the angle between the corresponding columns ci and cj in the communicability matrix: Tij = (cid:104)ci, cj(cid:105) (cid:107)ci(cid:107)(cid:107)cj(cid:107) , (4) being (cid:104) , (cid:105) the inner product and (cid:107)(cid:107) the vector norm. The definition of T depends uniquely on the topologi- cal constraints of the network encoded in the adjacency matrix A, plus the realistic assumption (embedded in C) that the influence or the information content decays with the length of the path. Despite being a purely topolog- ical measure, we recognized that, when the variance of the nodes is the same for all nodes, T corresponds ex- actly to the correlation matrix R of the Gaussian diffu- sion dynamics studied in Ref 41. Therefore, T formally closes the cycle for the search of a direct relation be- tween the structure of a network and the expected corre- lation pattern that this structure will tend to generate. Furthermore, this analytical measure unveils the funda- mental mechanism behind the contribution of network structure to the emerging function, which can be under- stood in terms of the similarity of influences that two nodes receive from the whole network via its complete path structure. Finally, we shall notice that T can be estimated for both directed and undirected graphs, as well as for weighted networks. In the case of undirected graphs C is symmetric but if the links are directed, then the columns of C determine the input profiles and the rows represent the profile of output influences of the nodes. It must be noted that despite the measure can be computed for any weighted adjacency matrix, it does not always make sense to do so. Because communicability is a measure of influence along the paths, it only has a direct physi- cal meaning when the weights of the links represent the coupling strength between the nodes, the flow capacity of the link or a compatible physical sense. Summarising, here we have introduced a topological estimator of a network's correlation structure. Although it ignores any specific dynamics of the nodes, it accounts for the fact that information or influence within a net- work propagates along all possible alternative paths and that naturally decays for longer paths. The measure ac- tually quantifies the similarity between the input profiles of nodes. In the following section we systematically in- vestigate how three fundamental features of a network (the weights of the links, the path length and the redun- dancy of paths) influence both the communicability and the topological similarity between nodes that, as shown above, can in fact be used as a proxy of their expected correlation. III. EFFECT OF TOPOLOGICAL PRIMITIVES In this section, we will investigate how three topolog- ical primitives we defined before, namely links' weights, graph distance and the presence of multiple alternative paths, sculpts the influence that one node exerts over an- other, setting aside the role of local nodes' dynamics. To this aim, we will focus on three simple classes of graphs: (i) chains, (ii) cycles and (iii) path-redundant networks. We will evaluate how manipulating critical parameters of these graphs leads to changes in the influence between given pairs of nodes a and b (measured by their commu- nicability, Cab) and in their topological similarity Tab. The contribution of link weight in the simplest case can be understood when analyzing how changes in the weight modulate the mutual influence of two nodes directly con- nected by a single link (see Supplemental Figure S1). As expected, incresing link weight is associated with increase in both the influence that one node exert over the other, as well as in their topological similarity. In general, A simple manner to obtain a better intuition of the behavior of both communicability and topological simi- larity is by studying how increasing the graph distance between two reference nodes affects both Cab and Tab in simple chain topologies. See top-left panels of Figure 1. it is possible to see how increasing the length of the path separating the two nodes correspond to a decrease in both their communicability and topolog- ical similarity, behavior directly determined by the de- cay formalized in the definition of the communicability. From the example, it is indeed evident how, for increasing lenghts of the chain, the input profiles of the two nodes at the ends of the chain (highlighted with green rectan- gles in the communicability matrices, top-left panels of Figure 1) become more and more antithetic, reflecting opposed whole-network influences. This behavior is cap- tured by the corresponding decrease in the nodes topo- logical similarity. The effect of uniformly varying the weights of all links in a chain is mainly quantitative: in fact, augmenting the weights in chains of fixed length in- creases the strength of the influence of each node over all other nodes in a way that is inversely proportional to the distance separating them, top-right panel of Fig- ure 1. See also Supplemental Figure S2 to appreciate the interaction of chain length and links' weight on C and T between the nodes at the two ends of a chain). 4 Due to the definition of communicability, the length of the path separating any two nodes as well as the weights of the links are the most important parameters defining the strength of their mutual influence and therefore their topological similarity as well. As such, chain topologies can be thought as a baseline to compare how more com- plex motifs such as cycles and path-redundant topolo- gies, which indeed incorporate chain topologies, modu- late both the resulting communicability and the topolog- ical similarity between given nodes. With this in mind, we compared Cab and Tab of the three model graphs (chains, cycles and path-redundant motifs) having corresponding graph diameter, i.e. having same longest path L. Bottom-left panel of Figure 1 pro- vides a schematic representation of different motifs hav- ing equivalent longest paths. For the case of chains and path-redundant topologies, we computed Cab and Tab for those two nodes that were more distant, in other words, those at the two extremes of (each) path, whereas for cyclic topologies, we always selected two adjacent nodes: this choice allowed us to clearly disentangle the contribu- tion that the direct link has on both the commmunicabil- ity and on the topological similarity, above and beyond the modulation produced by chains of different length. in chains both Cab and Tab decay as the distance between the extremal nodes increases, see bottom-right panels of Figure 1, blue lines. On the other hand the effect of the direct link is well illustrated for the case of cyclic architectures (bottom-right panels in Figure 1, red lines). In fact, the presence of a direct link importantly enhances and poses a lower bound for both Cab and Tab , that however exhibit a chain-like decay as the indirect path between them increases. As expected, This approach makes the difference between chains and cyclic primitives straightforward; however, in order to being able to properly understand how increasing path redundancy affects the communicability and hence topo- logical similarity, it can be of great help to directly quan- tify the difference between Cab and Tab calculated from chains and from redundant topologies for comparable path length. The two lowest-right panels in Figure 1 illustrate three examples of these differences, namely the cases with two (dark green lines), three (light green lines) and four (orange lines) redundant paths. From this anal- ysis, we appreciate that increasing the number of alterna- tive paths does increase both the total influence and the topological similarity between the two reference nodes at both ends of the paths, but that the magnitude of this increase decays with the length of the path, vanishing for paths longer than 5 links. Studying these simple networks (single link, chain, cy- cles and redundant paths) gives the opportunity to un- derstand the central relevance of the three topological primitives in shaping how the influences of nodes unfould through the graph. This information in turn determines how strong will be the expected correlation between any pair of nodes, approximated by their topological similar- ity. These simplified network models can thus be used 5 FIG. 1. Behavior of Communicability and Topological Similarity in simple network motifs. (Top-left) Communicability and Topological Similarity matrices of chains of different lengths. In the upper matrices, the red dots indicate the matrix entry corresponding to the communicability between the nodes at the two ends of the chains, whereas the green rectangles marks their whole-network input profiles (column vectors), that are used to calculate the topological similarity of the corresponding nodes (marked with the green dots in the lower matrices. (Top-Right) Communicability and Topological Similarity matrices of chains of constant length (L =21) for different links' weights, w. See top-left caption for the legend of red dots, green rectangles and green dots. (Bottom-left) Schematic representation of different graphs (chains, cycles and path-redundant architectures) having comparable longest path. The reference nodes for which both the communicability (Cab) and the topological similarity (Tab) where calculated are highlighted in yellow. (Bottom-Right) Upper panels: comparison of Cab and Tab of the three different graphs having comparable longest path. Line colors correspond to those in the schematic representation in the bottom-left region of the figure (Light blue lines: chains; Red lines: cycles; Dark green lines: two redundant paths; Light green lines: three redundant paths; Orange lines: four redundant paths). Lower panels: diffference between Cab and Tab of chains and redundant topologies. All results were obtained for constant links' weights and global coupling w = g = 1. to summarize some of the simplest forms of interactions sustained by these topological features. IV. UNDERSTANDING THE BRAIN'S SPONTANEOUS CORRELATION STRUCTURE In the previous section we have analyzed how simple network architectures could affect both communicability and topological similarity between a given pair of nodes. However, real networks are made of interwined assem- blies of those topological motifs that form intricate ar- chitectures and, together with the particular dynamical properties characterizing the system at hand, determine the emergence of complex patterns of interactions. In this section, we will try to find out) how much of the complex pattern of spontaneous correlations empirically observed in the resting brain can be explained just by the topology of the underlying anatomical structure. The main assumption we make is the following: if two brain regions receive similar influences from the entire network, then the probability that they will exhibit con- sistent co-activations is high; on the other hand, if the inputs they receive from the whole network are very dif- ferent, it is more likely that their covariance is weak. In Sections II and III, we showed how the the structure of a network can be used to estimate the strength of the influence that one node exerts over another trough the communicability C, and that C can be further used to quantify the topological similarity, T , of the input pro- files between any pair of nodes. We then demonstrated that T can in fact be interpreted as the expected correla- tion between the nodes in the network. As such, we will use T to estimate the unique contribution of topology to the empirical time-average correlation structure of the brain's spontaneous activity measured using resting-state fMRI. To this aim, we will compute the topological similar- ity matrix from the group-average structural connectiv- ity matrix (SC), and use it to estimate the correlation structure mirrored by the group-average empirical func- tional connectivity matrix (FC) obtained from resting- state fMRI. See Section VI for details about the em- pirical SC and FC matrices. The SC matrix stores the information about axonal pathways reconstructed using whole-brain diffusion tensor imaging and tractography, thus defines the whole-brain anatomical wiring diagram of the brain. It should be noted that, despite their re- producibility, current methods used to reconstruct fiber bundles from diffusion imaging are characterized by in- trinsic limitations constraining their accuracy: in fact, it is well-known that these reconstruction algorithms tend to favor the shortest, straightest and simplest path be- tween any two reference voxels42, which in turn impair their ability to accurately detect crossing fibers and long inter-hemispheric axons43,44. With this in mind, we fo- cused only on intra-hemipsheric structural and functional connectivity, in order to avoid the confounding effects of 6 FIG. 2. Effect of the global coupling. The figure illustrate the effect of the global coupling, g, onto the resulting topological similarity matrix, T . The three matrices have been obtained from the same empirical structural connectivity matrix, SC. accumulating errors due to the above mentioned limita- tions in sampling inter-hemispheric pathways. As mentioned in Section II, the communicability can be scaled by a factor g controlling the emerging collective dynamics, thus leading to the possibility of obtaining as many topological similarity matrices as the values of g used to scale C (see Figure 2 for three instances of T obtained for increasing values of g). For each hemisphere, we thus searched for the topo- logical similarity matrix T that best explained the ob- served intra-hemispheric correlation structure, by opti- mizing the communicability matrix according to a global scaling factor, g (Panel G and H in Figure 3). As a mesure of model fitting, we used the mean ab- solute error (MAE), an outlier-robust alternative of the mean squared error (MSE), a classic statistic to quantify the goodness of an estimator. From the scatter plots (Panels I,L of Figure 3) and the best-fitting T matrices (Panels M,N of Figure 3), it is possible to appreciate how properly accounting for the overall input pattern sustained by a given topology can indeed reduce the error in estimating the emergent correlations from the raw structural connectivity alone; in fact, in the example we pass from a E(SC,FC)≈ 0.42 to E(T ,FC)≈ 0.15). These results demonstrate that knowledge of the topol- ogy of whole-network input patterns of different brain re- gions, sustained by direct and indirect routes of multiple interweaved axonal bundles, can very much predict the time-average correlation structure observed from sponta- neous BOLD fluctuations, above and beyond the infor- mation about direct anatomical connections stored in the SC matrices. In the following, we will investigate how explicitly in- troducing local nodes' dynamics could affect the predic- tion of the average correlations. To this aim, we will make use of the Hopf normal model, that have success- fully used to predict mesoscopic brain activity45,46, and evaluate their ability to explain brain functional connec- tivity. By this comparison, we will be able to gain some insight about the contribution of adding local dynam- ics in determining the observed correlation structure, as well as to have a better understanding of the role of the 7 FIG. 3. Using topology to approximate the brain's spontaneous correlation structure. The figure shows results obtained separately for the left and right hemisphere. (A,B) Intra-hemispheric structural connectivity matrices (SC). (C,D) Scatterplot depicting the relationship between SC and the intra-hemispheric empirical correlation structure, summarized also in panels E and F. (E,F) Empirical functional connectivity matrices (FC) for the left and the right hemisphere. (G,H) Mean absolute error (MAE) between the empirical FC and the topological similarity T computed for different value of the global coupling parameter, g. For both hemispheres, minimal MAE (≈0.15) corresponded to g ≈1. The dotted lines correspond to the mean absolute error between the SC matrix and the empirical FC, which ≈0.45. (I,L) Scatterplots of the empirical FC and the best-fitting topological similarity for the two hemispheres. (M,N) Best-fitting Topological similarity matrices of the two hemispheres. underlying topology, shared either by the two models. A. Introducing local dynamics We will now explicitly introduce local dynamics to sim- ulate large-scale brain activity using the connectional ar- chitecture defined by the empirical SC matrix. By do- ing this, we will be able to estimate, through numeri- cal simulations, the correlation between different brain regions, and then compared these results with the aver- age empirical correlation matrix (Empirical FC) obtained from resting-state fMRI. Numerical simulations will be achieved through the use of the Hopf normal model, a dy- namical model able to reconcile noise-based approaches with models based on oscillators. This formalism is based on the normal form of a Hopf bifurcation45,47, a type of bifurcation that occurs when a system characterized by a stable fixed point loses its stability by exhibiting oscillations. As such, this model allows transitions be- tween asynchronous noise activity and oscillations, thus making it a good candidate to reproduce empirical data as observed either with EEG, MEG or fMRI45 -- 47. This model rely on the choice of two parameters, namely g, the global scaling of the strength of all the links in the network, and α, the parameter controlling the dynami- cal regime of each node. The bifurcation parameter will be set to 0, meaning that all nodes lie at the bifurca- 8 V. SUMMARY AND DISCUSSION The brain is a complex system and as such its overall dynamics cannot be fully understood without taking into account the rich patterns of interactions into which its components are inherently embedded into. The collective dynamics in a network results from the complex interplay between its underlying structure, the nature of the local dynamics of individual nodes, their specific working point and the manner in which they are coupled26,27,33. Here, we aimed at understand to what extent does the structure of a network drive the emerging patterns of interactions. Is it possible, knowing the complete wiring diagram of a network, to estimate its most likely correlation structure? Setting aside the large influence from the local dy- namics, we propose that the topological features which play major roles in shaping the interaction between two nodes are: (i) the strength of the links, (ii) the length of the path between the nodes, and (iii) the routing of information along multiple and redundant paths. A strong direct connection between two nodes is usually a reliable indicator of the strength of their functional interaction. In fact, in general direct links lead to more effective communications since the content of information (e.g., the amplitude of the perturbations) tends to decay over longer paths28. However, the flow of information from one node to another does not follow a unique path, but rather spreads along several. Therefore, the total influence that one node exerts over another is accumulated over all possible paths, of all lengths29 -- 31. The goal of this paper is that of answering the question of how the structure of a network contributes to sculpt the expected pattern of correlations that emerge when the network hosts a dynamical process. A direct appli- cation is that of understanding the extent to which the structural connectivity of the brain determines the large- scale correlation structure observed during rest. Accord- ingly, we have developed a graph measure, the topological similarity T , which estimates the expected correlation between two nodes based on the similarity of the "in- fluences" both receive from the whole network. If two nodes receive the same sets of inputs, then they should be strongly correlated. On the contrary, if they share no common inputs, their correlation tends to be weaker. Hence, this measure can be considered as a generaliza- tion of the matching index. In the analysis of graphs the matching index is typically used to quantify the number of common neighbours shared by two nodes. However, the matching index only accounts for the direct neigh- bours and ignores the rest of the complex fabric of inter- actions sustained by the entire network. To perform a complete comparison of the inputs accumulated over all paths, while accounting for the natural decay of signal power or information content in any real system, we con- sidered the columns of the communicabiilty measure32 as they directly correspond to the whole-network input profiles of the nodes. Although the actual decay rate of FIG. 4. Numerical simulations. The figure illustrate results for the two hemispheres obtained from numerical simulations using the Hopf normal model. (A,B) The model returned best results (MAE≈0.1) for values of the global coupling, g ≈ 0.12 for both hemispheres. As for data shown in Section IV, model fitting was performed using the mean absolute error (MAE). MAE between empirical structural (SC) and functional connectivity (FC) matrices is indicated by the dotted lines (≈ 0.42 in both hemispheres). (C,D) scatter plots of the empirical functional connectivity (FC) versus the simulated one obtained at the best-fitting global coupling. (E,F) In both panels, the upper triangles store the empirical FC values, whereas the lower triangles the corresponding ones obtained from simulations at the best fitting value of g. tion working point, region that has been demonstrated to give good approximate of the empirical resting-state FC46, and we will thus optimize for g. Results from nu- merical simulations are depicted in Figure 4. The in- troduction of complex local dynamics does indeed in- crease the capacity of predicting the time-average corre- lation structure (MAE(T )≈ 0.15, whereas MAE(Hopf)≈ 0.11)), even though the magnitude of this increase is rel- atively small, suggesting that most of the average struc- ture observed in correlated spontaneous BOLD fluctua- tions are in fact largely determined by the topology of the underlying network's architecture. the signals may differ across real systems, our choice of the communicability guarantees (due to its exponential decay) that the accumulation of perturbations over all possible paths, Eq. (1), converges for all adjacency ma- trices A. However, any other formalism implementing a decay of the influence between nodes as a function of their graph distance can be used to compute the topological simmilarity T , with the constrain that such formalism must be based on a set of coefficients assuring series con- vergence for any adjacency matrix. Recently, Zamora-L´opez et al. 41 found that consider- ing the communicability as the propagator kernel for the diffusion of Gaussian noise along the network allows to analytically estimate the time-average cross-correlation matrix R of the system. We realise that both approaches, the one starting from a dynamical system describing the propagation of perturbations41, and the other based uniquely on topological constrains and the assumption of exponential decay of influences, are indeed equivalent. In fact, we find that T ≡ R when the variance ξi of the Gaussian noise is the same for all nodes in the Gaus- sian diffusion system. Therefore we can conclude, with a high confidence, that T represents the most likely ex- pected correlation structure of a network only due to its underlying topology. As argued before, the other crucial ingredients for the collective dynamics on a network are the local dynamics of the nodes and the coupling functions, both of which can be either linear or nonlinear. The topological simi- larity T thus captures the tendency of the nodes of the network to correlate or to synchronise with each other. The contribution of the local dynamics and of the cou- pling function is to modulate this original background tendency set by the connection topology and either en- hance or disrupt the underlying patterns of correlation. In order to gain understanding on how each of the three topological features, namely, link weight, path length and path redundancy, precisely affect how influence propa- gates through the network, we have first investigated the behaviour of C and T between selected nodes in simple networks: chains and cycles of varying length, and path- redundant graphs. Not surprisingly, we have found that the strength of a direct link between two nodes is a major contributor to the intensity of their expected correlation. However, the presence of common inputs or redundant paths between them enhances the intensity of their in- teraction beyond the baseline determined by the strength of the link. If there is no direct link between two nodes, common inputs and redundant paths can trigger strong correlations between them. However, this influence tends to decay with the length of the paths. Although the pre- cise decay rate of the influence depends on the charac- teristics of the real system, it is worth noting that real networks with diameter larger than five are rather un- common. Specially in brain and neural networks which are dense. From a more general perspective, we shall no- tice that large efforts have been devoted in the literature to investigate which of the common network properties, 9 e.g., degree distribution, clustering coefficient, communi- ties or motifs, determine more prominently the collective dynamics on a network, particularly its capacity to syn- chronise48 -- 53. After our observations we can conclude that what truly matters for the collective dynamics is the path structure of the network which determines how far do perturbations reach and how their effect accumu- lates over redundant and recurrent paths. Thus, we fore- see that, to precisely understand how the typical network properties determine the collective dynamics, future work needs to identify how those properties alter the underly- ing network's path structure. Limitations and outlook The results we have here presented come with some limitations which shall be underlined. (A) The topologi- cal similarity T we have introduced represents an estima- tion of the time-averaged cross-correlation. Thus, it only estimates the spatial correlations between brain regions and does not capture temporal correlations. (B) Commu- nicability measure assumes an exponential decay of the influence with pathlength. This is not the only choice possible and other sets of coefficients kl exists which will lead the weighted version of the sum in Eq. (1) converge. The precise decay rate of influence, perturbations or in- formation in real systems will vary according to the sys- tem's nature. Thus, in some real applications it might be possible to identify the decay rate with pathlength and identify the "right" set of coefficients kl. As such, T is not restricted by the set of coefficients used in the com- municability measure proposed by32 .(C) We found here more convenient to quantify topological similarity as the cosine similarity of the input profiles, Eq. (4), as we found it more accurate in estimating the similarity between in- put profiles of small size, as in the case of short chains. However, other measures are possible, e.g., correlation or euclidean distance, without conceptually modifying the meaning of topological similarity. (D) for the analysis on the brain's connectivity and spontaneous correlations, we considered both hemispheres as if they were independent. The reason was to avoid biases due to the unreliability of tractography to identify inter-hemispheric fibers. Still, our comparisons are biased to some extent because the SC-based T and simulations consider both hemispheres as independent while the empirical resting-state measure- ments reflect the activity of the brain regions, which are, certainly, embedded on the whole network. Finally, we want to highlight the usefulness of both topological sim- ilarity T and the correlation matrix R of the Gaussian diffusion system41 to explore the functional connectiv- ity of synthetic and empirical networks. Because they estimate the expected correlation analytically, they are very fast to compute. For example, they allow to com- pare the effects of network perturbations, e.g., node or link lessions, without the need to run computationally expensive simulations. The only difference between T and R is that the formalism of R allows also to explore the effects of simulated inputs by increasing of decreasing the variance of the Gaussian noise at selected nodes. VI. MATERIALS AND METHODS To obtain the population average structural and func- tional connectomes twenty-one healthy volunteers (mean age 21.56 years; standard deviation 1.84 years; all males; all right handed) participated in five (5) resting-state and two (2) DTI scanning sessions. All participants signed in- formed consent to participate. The study was conducted at the School of Psychology, Birmingham and was ap- proved by the University of Birmingham Ethics Com- mittee. Data acquisition The scanning sessions were conducted at the Birm- ingham University Imaging Centre using a 3T Philips Achieva MRI scanner with a 32-channel SENSE head coil. T1-weighted anatomical data (175 slices; 1 × 1 × 1 mm3 resolution) were collected during the first ses- sion only. DTI data were collected in two sessions (23.3 2.5 days apart). The DTI acquisition consisted of 60 isotropically-distributed diffusion weighted direc- tions (b=1500 smm-2; TR=9.5s; TE=78ms; 75 slices; 2 × 2 × 2 mm3 resolution) plus a single volume without diffusion weighting (b = 0 smm-2, denoted as b0). The DTI sequence was repeated twice during each session, once following the Anterior-to-Posterior phase-encoding direction and once the Posterior-to-Anterior direction, to correct for susceptibility-induced geometric distortions54. Resting-state data were collected in five sessions (the first and the last collected in the same scanning session as the DTI data) using whole brain echo-planar imaging (EPI) (TR = 2s; TE = 35ms; 32 slices; 2.5 × 2.5 × 4 mm3 res- olution). Participants were instructed to have their eyes open and maintain fixation to a white dot presented at the centre of the screen. Whole-brain DTI tractography We 5.0.8 (FMRIB in the DTI processed data Software FSL version Library, http://fsl.fmrib.ox.ac.uk/fsl/fslwiki/) on a Red Hat Linux operating system. We corrected the data for susceptibility distortions, eddy currents and motion artifacts55. We subsequently rotated the gradient direc- tions (bvecs) to correct them for motion rotation42,56,57. We then used the corrected gradients in the Bayesian Estimation of Diffusion Parameters Obtained using Sampling Techniques (BedpostX) tool to generate a distribution model in each voxel58. We used the default parameters in BedpostX for the diffusion modelling: 2 10 fibers per voxel, weight of 1 for the secondary fibers, discard of the first 1000 iterations before sampling. We parcellated the brain into 116 areas using the Auto- mated Anatomical Labeling (AAL) atlas59. We followed a 4-step registration procedure to align the AAL atlas from MNI to native space: (a) align the non-weighted diffusion volume (b0) of each session to their midspace and create a midspace-template, (b) align the midspace- template to the anatomical (T1) scan, (c) align the T1 to the MNI template of FSL, and (d) invert and combine all the transformation matrices of the previous steps to ob- tain the MNI-to-native registration. The results of each step were visually inspected to ensure that the alignment was successful. Step (a) controls for potential bias to- wards the first session, when the T1 was acquired (similar methodology to Ref 60. The first two registrations are 6- dof linear transformations (rigid-body) since we aligned images of the same subject, while the third is 12-dof non- linear to warp the participants brain around the MNI template. The final matrix of step (d) was applied to the AAL atlas using nearest-neighbour interpolation to preserve the labels of the areas. We calculated the number of probabilistic streamlines starting from each AAL area and reaching any other AAL area by feeding the BedpostX model to the Probabilis- tic Tracking algorithm (ProbtrackX)61. The parameters we used in ProbtrackX are: 5000 samples per voxel, 2000 steps per sample until conversion, 0.5mm step length, 0.2 curvature threshold, 0.01 volume fraction threshold and loopcheck enabled to prevent streamlines from forming loops. We normalised the number of streamlines by the size of the seed area and thresholded streamlines lower than 1% of maximum (i.e. setting them to zero). We sub- sequently computed the undirected structural connectiv- ity matrix by averaging the normalised streamlines from area i to area j and from area j to area i. The results for each subject (in each DTI session) were organised into 42 weighted adjacency matrices A of size 116 × 116. Population-average structural connectome To estimate the population average structural connec- tivity (SC) we pooled the 42 SC matrices together (2 per subject) and considered only the reduced parcella- tion into 90 brain areas (45 per hemisphere) by exclud- ing the 26 regions of the cerebellum and the vermis. The 42 SC matrices contained a variable number L of undirected links ranging from L = 895 for the spars- est case (density ρ = 0.22) to L = 1279 for the dens- est (ρ = 0.32). We noticed that the simple average of the matrices into a single SC matrix by averaging the 42 values each link takes along the pool leads to an av- erage connectome with strongly biased network proper- ties. For example, this plain average SC matrix contained L = 1967 links, which is almost twice the number of links as in the individual matrices. In order to avoid this problem we have devised a method which automatically removes outlier links before performing the average. For ij} each link (i, j) we have initially a set of 42 weigths {ws where s = 1, 2, . . . , 42. The method searches for outlier weights (data-points falling out of 1.5 times the inter- quartile range) and removes them from the data pool. The search is iteratively repeated until no further outliers are detected and then the population-average SC weight for the link (i, j) is calculated as the average weight of the surviving values. In practice, the method converges very rapidly and it rarely performs more than 2 iterations per link. This method allows to clean the data without having to set an arbitrary hard threshold62 for the min- imally accepted prevalence of the link. Full details of the method are currently in preparation and will be pre- sented somewhere else. The resulting population-average SC matrix out of our iterative pruning method contains L = 1189 links (ρ = 0.30), which lies within the range of connectivity for the individual 42 matrices. For the simulations we treated the left and the right hemispheres independently, as two matrices of N = 45 ROIs because tractography is known to largely miss interhemispheric connections. Resting-state time-courses and functional connectome We the EPI resting-state pre-processed data in FSL version 5.0.8 (FMRIB Software Library, http://fsl.fmrib.ox.ac.uk/fsl/fslwiki/) on a Red Hat Linux operating system using MELODIC (Multivariate Exploratory Linear Optimized Decomposition into Independent Components). We corrected the data for motion and slice scan timing, removed the non-brain tissue, applied 5mm FWHM spatial smoothing and removed spike motion artifacts using WaveletDespike63. We subsequently applied high-pass temporal filtering and then extracted the average time-course from each AAL area. To estimate the population-average func- tional connectivity (FC) matrix we considered again only N = 90 brain areas excluding the cerebellum and the vermis. We concatenated the 105 sequences of resting-state signals (21 subjects, 5 sessions per subject) into a single long multivariate time-series and computed the Pearson correlation (z-Fisher corrected) for every pair of signals. The opposite procedure, to compute an FC matrix per session and averaging over the 105 FC matrices leads to almost identical results. Hopf normal model Within this model, the temporal evolution of the ac- tivity z of node j is given in the complex domain as: = [αj + iωj − z2] + σηj(t) dzj dt zj = ρjeiθj = xj + iyj (5) (6) 11 Where ω is the node's intrinsic frequency of oscilla- tion, α is the local bifurcation parameter (local because the model allows the possibility to assign a different value of α for each node in the network) and η is additive Gaus- sian noise with standard deviation σ. This system above shows a supercritical bifurcation at α = 0. Specifically, if αj is smaller than 0, then the local dynamic has a stable fixed point at zj = 0, while for αj values larger than 0 there exists a stable limit-cycle oscillation of frequency f = ω/2π . Whole-brain dynamics are described by the following coupled equations: = [αj − x2 j − y2 j ]xj − ωjyj + Cij(xi − xj) + σηxj(t) dxj dt dyj dt N(cid:88) i=1 N(cid:88) g g = [αj − x2 j − y2 j ]yj + ωjxj + Cij(yi − yj) + σηyj(t) (7) (8) i=1 Where Cij is the anatomical connectivity between nodes i and j, g is the global coupling factor and the standard deviation of gaussian noise is σ = 0.02. In this model the simulated activity corresponds to the BOLD signal of each node. The intrinsic frequency of each node was estimated as the peak frequency in the associ- ated narrowband (i.e., 0.04 - 0.07 Hz64) of the empirical BOLD signals of each brain region. We simulated, for ech of the two hemispherese (45 ROIs each), 330000 points using Euler's method for integration (dt = 0.001). The connectivity between all the regions of interest was de- fined using the empirical structural connectivity matrix (SC), and obtained timeseries were then used to compute the simulated correlation matrix (Simulated FC). ACKNOWLEDGMENTS We would like to thank Rui Wang and Caroline di Bernardi Luft for their help in collecting the data used in this study. This work was supported by (RGB) the FI-DGR scholarship of the Catalan Gov- ernment through the Ag`encia de Gesti´o d'Ajuts Uni- versitari i de Recerca, under agreement 2013FI-B1- 00099, (GZL) the European Union's Horizon 2020 re- search and innovation programme under grant agreement No. 720270 (HBP SGA1), (GD) the European Research Council Advanced Grant: DYSTRUCTURE (295129) and the Spanish Research Project PSI2013-42091-P, (ZK) European Community's Seventh Framework Pro- gramme [FP7/2007-2013] under agreement PITN-GA- 2011-290011, (VMK) European Community's Seventh Framework Programme [FP7/2007-2013] under agree- ment PITN-GA-2012-316746 and (MLK) by the Euro- pean Research Council Consolidator Grant: CAREGIV- ING (615539). 1B. Biswal, F. Zerrin Yetkin, V. M. Haughton, and J. S. Hyde, "Functional connectivity in the motor cortex of resting human brain using echo-planar MRI," Magn. Reson. Med. 34, 537 -- 541 (1995). 2A. Arieli, A. Sterkin, A. Grinvald, and A. Aertsen, "Dynamics of ongoing activity: explanation of the large variability in evoked cortical responses," Science 273, 1868 -- 1871 (1996). 3D. A. Gusnard and M. E. Raichle, "Searching for a baseline: Functional imaging and the resting human brain," Nat. Rev. Neurosci. 2, 685 -- 694 (2001). 4B. Mazoyer, L. Zago, E. Mellet, S. Bricogne, O. Etard, O. Houd´e, F. Crivello, M. Joliot, L. Petit, and N. Tzourio-Mazoyer, "Corti- cal networks for working memory and executive functions sustain the conscious resting state in man," Brain Res. Bull. 54, 287 -- 298 (2001). 5R. G. Shulman, D. L. Rothman, K. L. Behar, and F. Hyder, "Energetic basis of brain activity: implications for neuroimag- ing," Trends Neurosci. 27, 489 -- 495 (2004). 6M. D. Fox, A. Z. Snyder, J. L. Vincent, M. Corbetta, D. C. V. Essen, and M. E. Raichle, "The human brain is intrinsically or- ganized into dynamic, anticorrelated functional networks," Proc. Nat. Acad. Sci. USA 102, 9673 -- 9678 (2005). 7C. F. Beckmann, M. DeLuca, J. T. Devlin, and S. M. Smith, "Investigations into resting-state connectivity using independent component analysis," Phil. Trans. R. Soc. B 360, 1001 -- 1013 (2005). 8P. Fransson, "How default is the default mode of brain function?: Further evidence from intrinsic BOLD signal fluctuations," Neu- ropsychologia 44, 2836 -- 2845 (2006). 9S. M. Smith, P. T. Fox, K. L. Miller, D. C. Glahn, P. M. Fox, C. E. Mackay, N. Filippini, K. E. Watkins, R. Toro, A. R. Laird, and C. F. Beckmann, "Correspondence of the brain's functional architecture during activation and rest," Proc. Nat. Acad. Sci. USA 106, 13040 -- 13045 (2009). 10B. B. Biswal, M. Mennes, X.-N. Zuo, S. Gohel, C. Kelly, S. M. Smith, C. F. Beckmann, J. S. Adelstein, R. L. Buckner, S. Col- combe, A.-M. Dogonowski, M. Ernst, D. Fair, M. Hampson, M. J. Hoptman, J. S. Hyde, V. J. Kiviniemi, R. Kotter, S.-J. Li, C.- P. Lin, M. J. Lowe, C. Mackay, D. J. Madden, K. H. Madsen, D. S. Margulies, H. S. Mayberg, K. McMahon, C. S. Monk, S. H. Mostofsky, B. J. Nagel, J. J. Pekar, S. J. Peltier, S. E. Petersen, V. Riedl, S. A. R. B. Rombouts, B. Rypma, B. L. Schlaggar, S. Schmidt, R. D. Seidler, G. J. Siegle, C. Sorg, G.- J. Teng, J. Veijola, A. Villringer, M. Walter, L. Wang, X.-C. Weng, S. Whitfield-Gabrieli, P. Williamson, C. Windischberger, Y.-F. Zang, H.-Y. Zhang, F. X. Castellanos, and M. P. Milham, "Toward discovery science of human brain function," Proc. Nat. Acad. Sci. USA 107, 4734 -- 4739 (2010). 11H. Laufs, K. Krakow, P. Sterzer, E. Eger, A. Beyerle, A. Salek- Haddadi, and A. Kleinschmidt, "Electroencephalographic signa- tures of attentional and cognitive default modes in spontaneous brain activity fluctuations at rest," Proc. Nat. Acad. Sci. USA 100, 11053 -- 11058 (2003). 12M. J. Brookes, M. Woolrich, H. Luckhoo, D. Price, J. R. Hale, M. C. Stephenson, G. R. Barnes, S. M. Smith, and P. G. Mor- ris, "Investigating the electrophysiological basis of resting state networks using magnetoencephalography," Proc. Nat. Acad. Sci. USA 108, 16783 -- 16788 (2011). 13A. Ponce-Alvarez, G. Deco, P. Hagmann, G. L. Romani, D. Man- tini, and M. Corbetta, "Resting-State Temporal Synchronization Networks Emerge from Connectivity Topology and Heterogene- ity," PLOS Comput Biol 11, e1004100 (2015). 14S. D. Keilholz, J. C. W. Billings, K. Wang, A. Abbas, C. Hafe- neger, W. J. Pan, S. Shakil, and M. Nezafati, "Multiscale net- work activity in resting state fMRI," in 2016 38th Annual Inter- national Conference of the IEEE Engineering in Medicine and Biology Society (EMBC) (2016) pp. 61 -- 64. 12 15G. Deco, P. Hagmann, A. Hudetz, and G. Tononi, "Modeling resting-state functional networks when the cortex falls asleep: Local and global changes," Cereb. Cortex 24, 3180 -- 3194 (2013). and J. Binder, "Spin-glass model predicts metastable brain states that diminish in anesthesia," Front. Syst. Neurosci. 8, 234 (2014). 16A. Hudetz, C. Humphries, 17A. Hudetz, X. Liu, and S. Pillay, "Dynamic repertoire of intrin- sic brain states is reduced in propofol-induced unconsciousness," Brain Connectivity 0, 1 -- 13 (2014). 18R. G. Bettinardi, N. Tort-Colet, M. Ruiz-Mejias, M. V. Sanchez- Vives, and G. Deco, "Gradual emergence of spontaneous corre- lated brain activity during fading of general anesthesia in rats: Evidences from fMRI and local field potentials," NeuroImage 114, 185 -- 198 (2015). 19C. S. Zhou, L. Zemanov´a, G. Zamora-L´opez, C.-C. Hilgetag, and J. Kurths, "Hierarchical organization unveiled by functional connectivity in complex brain networks," Phys. Rev. Lett. 97, 238103 (2006). 20C. J. Honey, R. Kotter, M. Breakspear, and O. Sporns, "Net- work structure of cerebral cortex shapes functional connectivity on multiple time scales," Proc. Nat. Acad. Sci. USA 104, 10240 -- 10245 (2007). 21J. G´omez-Gardenes, G. Zamora-L´opez, Y. Moreno, and A. Are- nas, "From modular to centralized organization of synchroniza- tion in functional areas of the cat cerebral cortex," PLoS ONE 5, e12313 (2010). 22G. Deco, V. Jirsa, A. R. McIntosh, O. Sporns, and R. Kotter, "Key role of coupling, delay, and noise in resting brain fluctua- tions," Proc. Nat. Acad. Sci. USA 106, 10302 -- 10307 (2009). 23G. Deco, V. K. Jirsa, and A. R. McIntosh, "Emerging concepts for the dynamical organization of resting-state activity in the brain," Nat. Rev. Neurosci. 12, 43 -- 56 (2011). 24G. Deco, V. K. Jirsa, and A. R. McIntosh, "Resting brains never rest: computational insights into potential cognitive ar- chitectures," Trends Neurosci. 36, 268 -- 274 (2013). 25G. Deco and M. L. Kringelbach, "Great expectations: Using whole-brain computational connectomics for understanding neu- ropsychiatric disorders," Neuron 84, 892 -- 905 (2014). 26A. Mess´e, D. Rudrauf, H. Benali, and G. Marrelec, "Relating structure and function in the human brain: relative contributions of anatomy, stationary dynamics, and non-stationarities," PLoS Comput. Biol. 10, e1003530 (2014). 27G. Schmidt, G. Zamora-L´opez, C. Zhou, and J. Kurths, "Sim- ulation of large scale cortical networks by individual neuron dy- namics," Int. J. Bif. & Chaos 20, 859 -- 867 (2010). 28V. Latora and M. Marchiori, "Efficient behavior of small-world networks," Phys. Rev. Lett. 87, 198701 (2001). 29B. A. Huberman and L. A. Adamic, "Information dynamics in the networked world," in Complex Networks, Lecture Notes in Physics No. 650, edited by E. Ben-Naim, H. Frauenfelder, and Z. Toroczkai (Springer Berlin Heidelberg, 2004) pp. 371 -- 398. 30D. J. Ashton, T. C. Jarrett, and N. F. Johnson, "Effect of conges- tion costs on shortest paths through complex networks," Phys. Rev. Lett. 94, 058701 (2005). 31A. Trusina, M. Rosvall, and K. Sneppen, "Communication boundaries in networks," Phys. Rev. Lett. 94, 238701 (2005). 32E. Estrada and N. Hatano, "Communicability in complex net- works," Phys. Rev. E 77, 036111 (2008). 33L. Huang, Q. Chen, Y.-C. Lai, and L. Pecora, "Generic be- haviour of master-stability functions in coupled nonlinear dy- namical sstems," Phys. Rev. E 80, 036204 (2009). 34S. P. Borgatti, "Centrality and network flow," Soc. Networks 27, 55 -- 71 (2005). 35V. Colizza, A. Flammini, M. A. Serrano, and A. Vespignani, "Detecting rich-club ordering in complex networks," Nat. Phys. 2, 110 -- 115 (2006). 36J. Goni, M. P. v. d. Heuvel, A. Avena-Koenigsberger, N. V. d. Mendizabal, R. F. Betzel, A. Griffa, P. Hagmann, B. Corominas- Murtra, J.-P. Thiran, and O. Sporns, "Resting-brain functional connectivity predicted by analytic measures of network commu- 13 tainty in diffusion weighted mr imaging." Magn. Reson. Med. 72, 227 -- 236 (2003). 59N. Tzourio-Mazoyera, B. Landeaub, D. Papathanassioua, F. Crivelloa, O. Etarda, N. Delcroixa, B. Mazoyerc, and M. Jo- liot, "Automated anatomical labeling of activations in SPM using a macroscopic anatomical parcellation of the MNI MRI single- subject brain," Neuroimage 1 (200215). 60S. M. Smith, N. De Stefano, M. Jenkinson, and P. M. Matthews, "Normalized accurate measurement of longitudinal brain change." J. Comput. Assist. Tomogr. 25, 466 -- 475 (2001). 61T. Behrens, H. Berg, S. Jbabdi, M. Rushworth, and M. Woolrich, "Probabilistic diffusion tractography with multiple fibre orienta- tions: What can we gain?" Neuroimage 34 (2007). 62M. A. de Reus and M. van den Heuvel, "Estimating false positives and negatives in brain networks," NeuroImage 70, 402 -- 409 (2013). 63A. X. Patel, P. Kundu, M. Rubinov, P. S. Jones, P. E. V´erts, K. D. Ersche, J. Suckling, and E. T. Bullmore, "A wavelet method for modeling and despiking motion artifacts from resting- state fmri time series." NeuroImage 95, 287 -- 304 (2014). 64E. Glerean, J. Salmi, J. M. Lahnakoski, I. P. Jaaskelainen, and M. Sams, "Functional magnetic resonance imaging phase syn- chronization as a measure of dynamic functional connectivity," Brain connectivity 2, 91 -- 101 (2012). nication," Proc. Nat. Acad. Sci. USA 111, 833 -- 838 (2014). 37A. Avena-Koenigsberger, B. Misi´c, R. Hawkins, A. Griffa, P. H. andJoaqu´ın Goni, and O. Sporns, "Path ensembles and a tradeoff between communication efficiency and resilience in the human connectome," Brain Struct. Funct. (2016). 38F. Harary and A. Schwenk, "The spectral approach to determin- ing the number of walks in a graph," Pacific Journal of Mathe- matics 80, 443 -- 449 (1979). 39J. Bang-Jensen and G. Z. Gutin, Digraphs: Theory, Algorithms and Applications (Springer, 2008). 40E. Estrada, N. Hatano, and M. Benzi, "The physics of communi- cability in complex networks," Phys. Reps. 514, 89 -- 119 (2012). 41G. Zamora-L´opez, Y. Chen, G. Deco, M. L. Kringelbach, and C. Zhou, "Functional complexity emerging from anatomical con- straints in the brain: the significance of network modularity and rich-clubs," Scientific Reports 6, 38424 (2016). 42D. K. Jones and M. Cercignani, "Twenty-five pitfalls in the anal- ysis of diffusion mri data." NMR Biomed. 23, 803 -- 820 (2010). 43C. Thomas, Q. Y. Frank, M. O. Irfanoglu, P. Modi, K. S. Saleem, D. A. Leopold, and C. Pierpaoli, "Anatomical accuracy of brain connections derived from diffusion mri tractography is inherently limited," Proceedings of the National Academy of Sciences 111, 16574 -- 16579 (2014). 44S. Jbabdi, S. N. Sotiropoulos, S. N. Haber, D. C. Van Essen, and T. E. Behrens, "Measuring macroscopic brain connections in vivo," Nature neuroscience 18, 1546 -- 1555 (2015). 45F. Freyer, J. A. Roberts, R. Becker, P. A. Robinson, P. Ritter, and M. Breakspear, "Biophysical mechanisms of multistability in resting-state cortical rhythms," The Journal of Neuroscience 31, 6353 -- 6361 (2011). 46G. Deco and M. L. Kringelbach, "Metastability and coherence: extending the communication through coherence hypothesis us- ing a whole-brain computational perspective," Trends in neuro- sciences 39, 125 -- 135 (2016). 47F. Freyer, J. A. Roberts, P. Ritter, and M. Breakspear, "A canonical model of multistability and scale-invariance in biologi- cal systems," PLoS Comput Biol 8, e1002634 (2012). 48A. E. Motter, C. S. Zhou, and J. Kurths, "Enhancing complex- network synchronization," Europhys. Lett. 69, 334 -- 340 (2005). 49M. di Bernardo, F. Garofalo, and F. Sorrentino, "Effects of de- gree correlation on the synchronizability of networks of nonlinear oscillators," (Proc. 44th IEEE Conference on Decision and Con- trol, and the European Control Conference WeA14.1, 2005). 50F. Atay, T. Biyikoglu, and J. Jost, "Network synchronization: Spectral versus statistical properties," Physica D 224, 35 -- 41 (2006). 51C. Zhou and J. Kurths, "Hierarchical synchronization in networks of oscillators with heterogeneous degrees," Chaos 16, 015104 (2006). 52X. Wu, B. Wang, T. Zhou, W. Wang, M. Zhao, and H. Yang, "Synchronizability of highly clustered scale-free networks," Chi- nese Phys. Lett. 23 (4), 1046 -- 1049 (2006). 53A. Arenas, A. D´ıaz-Guilera, J. Kurths, Y. Moreno, and C. S. Zhou, "Synchronization in complex networks," Phys. Reps. 469, 93 -- 153 (2008). 54J. L. Anderson, S. Skare, and J. Ashburner, "How to correct susceptibility distortions in spin echo-panar images: application to diffusion tensor imaging." NeuroImage 20, 870 -- 888 (2003). 55J. Anderson and S. Sotiropoulos, "An integrated approach to correction for off-resonance effects and subject movement in dif- fusion mr imaging." NeuroImage 125, 1063 -- 1078 (2016). 56A. Leemans and D. Jones, "The b-matrix must be rotated when correcting for subject motion in dti data." Magn. Reson Med 61 (2009). 57A. Ersiz, V. E. Arpinar, S. Dreyer, and L. T. Muftuler, "Quanti- tative analysis of the efficacy of gradient table correction on im- proving the accuracy of fiber tractography." Magn. Reson. Med. 72, 227 -- 236 (2014). 58T. E. Behrens, H. J. Berg, S. Jbabdi, M. F. S. Rushworth, and M. W. Woolrich, "Charaterization and propagation of uncer- "How structure sculpts function: unveiling the contribution of anatomical connectivity to the brain's spontaneous correlation structure." by Supplementary material for: Ruggero G. Bettinardi, Gustavo Deco, Vasilis M. Karlaftis, Timothy J. Van Hartevelt, Henrique M. Fernandes, Zoe Kourtzi, Morten L. Kringelbach and Gorka Zamora-L´opez 1 2 FIG. S1. Effect of Single Link Weight. (A) graph and matrix representation of two nodes connected by a direct link. (B) Communicabity between the two nodes, Cab, for diffferent values of link's weight. (C) Topological Similarity between the two nodes, Tab, for diffferent values of link's weight. On top of the panel are exemplified the communicability matrices obtained for different values of link's weight. Remember that Tab is computed as the cosine similarity between the two columns of the communicability matrix. 3 FIG. S2. Effect of Link Weight in chain and cyclic topologies. (A) Communicability Cab between the two nodes a and b at the ends of chains of different length for varying links' weights. (B) Topological similarity Tab between the two nodes a and b at the ends of chains of different length for varying links' weights. The blue line in Panels A and B correspond to the links' weight value used in chain topologies in the main text of the paper (w =1.) (C) Communicability Cab between two adjacent nodes a and b for cycles of increasing perimeter (N , number of nodes in the cycle) for varying links' weights. (D) Topological similarity Tab between two adjacent nodes a and b for cycles of increasing perimeter (N , number of nodes in the cycle) for varying links' weights. The red line in Panels C and D correspond to the links' weight value used in cyclic topologies in the main text of the paper (w =1.)
0910.2741
2
0910
2010-04-07T22:23:55
Fractals in the Nervous System: conceptual Implications for Theoretical Neuroscience
[ "q-bio.NC" ]
This essay is presented with two principal objectives in mind: first, to document the prevalence of fractals at all levels of the nervous system, giving credence to the notion of their functional relevance; and second, to draw attention to the as yet still unresolved issues of the detailed relationships among power law scaling, self-similarity, and self-organized criticality. As regards criticality, I will document that it has become a pivotal reference point in Neurodynamics. Furthermore, I will emphasize the not yet fully appreciated significance of allometric control processes. For dynamic fractals, I will assemble reasons for attributing to them the capacity to adapt task execution to contextual changes across a range of scales. The final Section consists of general reflections on the implications of the reviewed data, and identifies what appear to be issues of fundamental importance for future research in the rapidly evolving topic of this review.
q-bio.NC
q-bio
Fractals in the Nervous System: conceptual implications for Theoretical Neuroscience. Gerhard Werner [email protected] Department of Biomedical Engineering University of Texas at Austin, TX. Abstract This essay is presented with two principal objectives in mind: first, to document the prevalence of fractals at all levels of the nervous system, giving credence to the notion of their functional relevance; and second, to draw attention to the as yet still unresolved issues of the detailed relationships among power law scaling, self-similarity, and self-organized criticality. As regards criticality, I will document that it has become a pivotal reference point in Neurodynamics. Furthermore, I will emphasize the not yet fully appreciated significance of allometric control processes. For dynamic fractals, I will assemble reasons for attributing to them the capacity to adapt task execution to contextual changes across a range of scales . The final Section consists of general reflections on the implications of the reviewed data, and identifies what appear to be issues of fundamental importance for future research in the rapidly evolving topic of this review. Contents 1. Introduction 2. Power-law scaling in neuronal structures and processes 2.1 Neuronal Morphology 2.2 Peripheral nervous system 2.2.1 Ion channels 2.2.2 Point process analysis in peripheral nerves and neurons 2.3 The mesoscopic level 2.4 The macroscopic level 2.4.1 Fractals in Brain Networks 2.4.2 Fractals and Criticality of Brain States 2.4.3 Significance of Brain Criticality 3. Psychological and Behavioral Processes 3.1 Symbol processing and Fractals 3.2 Motor Behavior and Allometric control processes 4. Processes that generate power-law distributions 5. Fractals in Action 5.1 Complexity Matching Effect 5.2 Linking across many scales of space and time 6. Summary and final thoughts 7. References 1. Introduction Fractals, introduced by Mandelbrot in 1977, are in the spatial domain considered to be self-similar geometric objects with features on an infinite number of scales. In the analysis of time series, fractal time describes highly intermittent self-similar temporal behavior that does not possess a characteristic time scale. Their statistical analysis can provide access to understanding the dynamics of complex systems. Not possessing a single characteristic scale, static and dynamical fractals, measured on different scales of space and time, respectively, can be characterized by power functions whose (usually non-integer) exponents are their fractal Dimensions. In this essay, principal emphasis is on random fractals which include a stochastic element in their generation. Fractals are signals which display scale-invariant, self-similar behavior; they can be analyzed by decomposing a signal into a hierarchy of temporal and spatial scales that may cover the wide range between coarse-scale long-term and high-frequency fine-scale fluctuations. If the relationship of the property under consideration is simple with respect to change of scale, the process is considered monofractal , the process can then be characterized by a single scaling exponent which is related to the fractal dimension or the spectral exponent of the process. It can be expressed as the Hurst exponent of the process (Mandelbrot, 1968; Koutsoyiannis, 2002). However, in some instances the scaling behavior may not be adequately characterized by a single, stationary scaling exponent. In such cases, several scaling exponents may be required, each exponent locally pertaining to some portion of the data stream. Such multifractal signals are represented by the histogram of the Hoelder exponents, also known as the singularity spectrum (Muzy et al., 1993). Power law scaling and other manifestations of fractal and self-similar patterns in space and/or time can be identified at all levels of neural organization. With few exceptions, these observations remained largely islands in the otherwise rapidly advancing theoretical Neuroscience with different priorities. However, recent advances in methodology of measurement of fractal connectivity at higher levels of brain organization have led to a proliferation of new data. This now calls for integrating fractality with other insights into brain organization and complexity, notably in the light of the substantial evidence for the brain being a complex system in a regime of criticality, as understood in statistical physics (Chialvo,2004, 2008; Kitzbichler et al, 2009; Fraiman et al, 2009; Werner, 2007b, 2009a,b). Like in other physiological systems manifesting fractal patterns (see for instance: Bassigthwaighte et al, 1994; West and Deering, 1995; Iannacone and Khoka, 1995; West, 2006) the question of ubiquity of power law scaling needs to be addressed in relation to other features of brain organization. Similarly, is there a relation between fractal organization and the propensity for phase transitions of critical systems? Is there a bridge between coarse graining (including renormalization group transformation) and fractality ? And, most importantly, can fractal properties be viewed as playing a role for the functional integration among different levels of neuronal organization as Andersen (2000) suggests in an article entitled “From Molecules to Mindfulness”. Generalizing from a comprehensive theory of organization and interactions at the molecular level, Agnati et al (2004, 2009) view the Central Nervous System as a nested network at all levels of organization, in the image of the Russian Matryoshka dolls: self-similar structures being embedded within one another. This theory is elaborated in great detail, with a “Fractal (self-similarity) Logic” operating on a set of identical rules which would govern the relation between successive levels of the nested system. Giesinger’s ( 2001 ) comprehensive overview of scale invariance in Biology provides the background of this review, as do the insights gained in Physics through the work of Wilson (1979 ) and Kadanoff (1990), amongst many others. In addition, Turcotte (1999) discussed at great length the relation between aspects of self-organized criticality and fractal scaling from the points of view of “avalanche” behavior and systematic properties of correlation length, specifically directing attention to inverse cascade and site-percolation models of the well- known forest fire paradigm. While none of the issues discussed in the following will receive a definitive answer, I will aim at an explicit formulation of the network of interrelated factors that constitute the territory in which new perspectives and potential solutions may lie. With the agenda set forth in the foregoing, the organization of the presentation is as follows: I will first briefly review the neuroscience literature on fractals, organized by level of neuronal organization, from ion channels to cortical networks , and to psychological and behavioral functions. Criticality of brain states, Allometric control and the origin of fractals by phase transitions of complex systems will be addressed in more detail. This will be followed by a brief overview of the essentials features of the theory of fractal generators, including random walk theory and fractional differential operators. Having laid out the background in this manner, I will consider relations between renormalization group transformation and fractals as having some potential bearing on the apparent ubiquity and universality of power law scaling in neural structures and processes, and its relation to criticality. Finally, I will direct attention to the amazing consequence of self-similarity which assures the telescoping of different levels of structural and functional organization to constitute a fractal object or time series. This will lead me to posing the ultimate question: is there a process for unpacking interactions between different levels of the fractal object, responsive to circumstances and conditions, which eludes us entirely? If it existed, fractals would surely be a most extraordinary design principle for operational economy in complex systems. 2. Power-law scaling in neuronal structures and processes. This section is intended to summarize essential aspects of fractal properties at each of the conventionally designated organizational levels, as the basis for conceptual consideration of relations across these levels. However, a word of caution is in order: the sketches of observational data in this section encompass a vast variety of biological substrates, conditions of observation, and methods of measurement. This heterogeneity imposes limits on generalizations, as do the differences of criteria for identifying fractal or self -similar features in the data. Potential pitfalls were discussed and illustrated in LaBarbera’s (1989) useful (largely pedagogic) publication. More recently, Eke et al (2002), Deligniers et al (2006) and Clauset et al (2009) set forth stringent criteria for design, collection and interpretation of data for identifying and categorizing fractal properties. The latter authors are very specific in formulating a principled statistical framework, combining maximum likelihood fitting methods with goodness of fit tests; they demonstrate examples of data that had been conjectured to represent power law fits, but did not withstand the rigor of their tests. Touboul and Destexhe (2009) also suggest that apparent power law scaling may in some instances not be supported by more stringent statistical criteria. Of particular relevance to the topic of Section 2.3 is their claim that experimentally observed power law scaling must not considered proof of self- organized criticality, lest there be other supporting evidence available. The analysis, synthesis and estimation of fractal-rate stochastic point processes is reviewed and illustrated with examples by Thurner et al. (1997). Note also that the distinction between mono- and multifractal scaling is sometimes difficult to draw (Kadanoff et al, 1989). Conceivably, some of the variations among the reports reviewed in subsequent sections may be attributable to procedural differences among studies; other reports may not meet the rigorous statistical criteria of Clauset et al. Nevertheless, I submit that the majority of experimental data on fractals in neural structures give collectively adequate reason for ascribing to them wide-spread functional significance. At least the results based on wavelet analysis appear immune to methodological criticism (see Section 2.4). 2.1 Neuronal morphology In the foundational work “The fractal geometry of Nature”, Mandelbrot (1977) wrote “it would be nice if neurons - he mentioned specifically Purkinje cells in the cerebellum- turned out to be fractal”: Nature obliged abundantly as the following sample of findings with dendrites, neuron cell bodies and glia cells indicates. Studying the branching pattern of dendritic trees of retina neurons, Caserta et al (1990) identify by box counting fractal shapes with a fractal dimension of approximately 1.7, which can be explained by a diffusion limited aggregation model (Witten and Sander, 1981); but fractal dimension measured by different methods (for instance comparing box counting with cumulative mass method) gives appreciably different values (Caserta et al, 1955). A fractal structure was observed by Kniffki et al (1994 ) for the branching dendrite patterns of thalamic neurons in Golgi impregnated specimens. In a separate series, a scaling relation for bifurcations within the dendrite trees was ascertained (Kniffki et al, 1993). Significant species differences in fractal dimensions of dendrite arborizations in dorsal horn spinal cord neurons (Milosevic et al, 2007) may be attributable to species differences in peripheral somesthetic sensibility (the dorsal horn neurons being the first receiving station of this type of afferent input). Fractal analysis also reveals a distinct differentiation of neuron types in the different laminae of the dorsal horn (Milosevic et al, 2005). Differences in regional connectivity and functional capacity amongst different regions in visual cortex pyramidal neurons are also associated with marked variation in the fractal dendrite branching structure (Zietsch and Elston, 2005). Fractal analyses provide a measure of space filling of dendrite arbors which, in a study by Jelinek and Elston (2001) , differentiates in the macaque visual cortex the two known processing streams between primary and secondary visual area by differences in fractal properties. These investigators had undertaken a meticulous examination of criteria for ‘quality control’ in studies of this nature, from the stage of pre- processing of tissue specimens to comparative evaluation of methods for determining fractal dimension (Jelinek et al, 1995).Examining the connectivity repertoire of basal dendrite arbors of pyramidal neurons, Wen et al (2009) determined a universal power law scaling for dendrite length and radius, suggesting that the dendrite arbors are constructed by statistically similar processes; moreover, fragments of an arbor are statistically similar to the entire arbor, thus displaying self-similarity. These design features are thought to maximize functionality for a fixed dendrite cost. Additional evidence comes from digital image analysis which enabled Smith et al (1989) to determine the fractal dimension of neuron contours. Results obtained with conventional methods of scaling analysis are corroborated by Wavelet Packet fractal analysis (Jones and Jelinek, 1989). Multifractals were identified for cortical pyramidal cells while, in comparison, neurons of synRas transgenic mice display less complex arborization patterns (Schierwagen, 2008.). Shape complexity of neurons and elements of microglia in human brain can be classified over a range for fractal dimensions which is different for normal and pathological brains (Karperien et al, 2008 ). The sequence of developmental stages of oligodendrocytes, tracked the basis of their immunoreactivity, parallels changes in fractal dimension (Bernard et al, 2001). Fractal analysis of cell ramification and space filling patterns differentiate microglia cells into two categories, depending on whether their fractal dimension did, or did not increase after brain injury (Soltys et al, 2001). The scaling law for the cortical magnification factor in primate visual cortex is an illustrative example with functional significance: as is well known, the part of the visual scene corresponding to the eye center is represented densely at the cortex, becoming progressively sparser towards the periphery. It turns out that the scaling law for the sampling density away from area centralis is a power function which assures locating a peripheral target in the shortest time (Koulakov, 2010). As brain size increases, the cortex thickens only slightly, but the degree of sulcal convolution increases dramatically, indicating that human cortices are not simply scaled versions of one another (Im et al., 2008). Changizi (2003) infers several organizing principles of Neocortex from scaling relations among its components: e.g. diameters of neural structures predispose for efficient transport through neuron arborizations; economical wiring reflects well-connectedness within given volumes of neural tissue. These relations are viewed to represent a universal law for scaling that applies to hierarchical complexity and combinatorial systems, generally (Changizi, 2001 b). As one among several instances of scaling in the cerebral cortex, Changizi (2001 a) also shows that axon cross sectional area increases in Phylogeny with brain size, presumably compensating increase of conduction distances with conduction velocity (see Section 3.1). A synthesis of comparative neuroanatomy with biophysics leads Harrison et al (2002) to conclude that scaling trends in morphological specializations at the cellular level may constitute functional adaptations. One of the examples in support of their thesis is the role of component scaling for managing the conflicting developmental trends of increasing brain size and surface folding on the one hand, and the requirement for optimizing energy requirements and processing speed, on the other. The role of scaling relations for brain growth becomes evident when its scaling relation is disrupted by preterm birth in a dose-dependent, sexually dimorphic fashion ; it directly parallels the incidence of neuro-developmental impairments in preterm infants (Kapellou et al., 2006). Taken together, the observations surveyed in the foregoing two paragraphs suggest that fractal dimension of neuronal and glia elements bear some relations to developmental, functional and pathological conditions of neural tissue. This warrants a few conceptual considerations: Bieberich (2002) attaches neural-computational significance to the self- similarity of dendritic branching as a platform for economical information compression and recursive algorithms. On the same self-similarity principle, Pellionisz (1989) envisages a fractal growth model of dendritic arbors by iterated code repetition as process for global construction of fractals (see for instance: Barnsley & Dempko, 1985): the essential underlying theme is to both reduce complexity of generating, and at the same time conserving the full richness of the dendrite arbor. I will expand on this principle in later section of this essay. Among the not yet explored implications of dendrite fractal arborizations are the effect they may induce on the dynamics of processes and critical phenomena in dendrite spines for which they are a supporting platform: In Statistical Physics, such effects obtain when the neighborhood relations among interacting elements (for instance: Ising spins or coupled maps (Cosenza and Kapral, 1992) are themselves provided by a self-similar fractal lattices, such as the Sierpinsky Gasket (Gefen et al 1980), rather than an Euclidean geometric base. In an extension of fractal analysis to features of complex neural structures, Zhang (2006) determined the Magnetic Resonance image-based fractal dimension of white matter of human brain. This method was shown to accurately quantify white matter structural complexity in three dimensions, and detect age-related degenerative changes. Tractography based on Diffusion Tensor Imaging enabled Katsaloulis and Vergenelakis (2009) to determine fractal dimension, self-similarity and lacunarity of neuron tracts in human brain. The lacunarity analysis is understood to indicate the distribution of fractal neuron tracts of different length scales, as evidence of connections between different neuron ensembles. Another extraordinary technical advance made it possible to determine the fractal properties of receptor density and distribution in human brain, using Positron Emission Tomography (PET) and Single -photon Emission Tomography (SPET) (Kuikka and Tiihonen, 1998). 2.2 The peripheral nervous system: ion channels, point process analysis of activity in peripheral nerves and individual neurons 2.2.1: Ion Channels. Turning to primarily functional aspects of fractality in neural systems, attention focuses in this section on temporal aspects of ion channel gating and its relation to time series of neuronal discharge patterns. The following collage of data obtained with different experimental conditions as well as modeling studies consistently supports the dominant presence of fractal features in the functional manifestations at the levels under consideration. The kinetics of Ion transport across neuronal membranes occurs, in part, via ion channels . Application of the Patch clamp technique made it possible to follow the time course of channel opening and closing precisely. Typically, the rate of channel opening and closing opening fluctuates, changing at times suddenly from periods of great to periods of slow activity. This pattern served as clue to surmise an underlying fractal process with infinite variance. On this basis Liebowitch et al (1987; 2001) asked how the switching probabilities at one time scale of observation are related to those at another time scale. It turned out that these probabilities (defined as effective kinetic rate) at a given time scale are characterized by fractal scaling, and that effective kinetic rates for different time scales of observation display self -similarity: there are bursts within bursts of openings and closings. The suggestion is that energy barriers in stochastically switching protein conformational states are the underlying mechanism (for a detailed account, see Ch. 8 in Bassingthwhaite et al, 1994). A different version that also accounts for the power law relationship of ion channel gating kinetics assumes that ion channel proteins have a very large number of states, all of similar energy, making the gating process more akin to a diffusion (Millhauser 1988). Recent theoretical modeling defined more precisely the conditions that give rise to the power law distributions in relation to the activation barriers, compatible with the known Physics of proteins (Goychuck and Hanggi, 2002). Roncaglia et al (1993) developed on theoretical grounds a stringent criterion for ascertaining the validity of the fractal theory by evaluating the experimental distribution of channel closing times in terms of the Hurst phenomenon. A few years thereafter, Varanda at al. (2000) delivered the evidence for Ca-activated K channels in the form of long term correlations of open and closed dwell times, expressed as Hurst coefficients of the order of 0.6, which alternative Markovian models failed to satisfy. Before proceeding to discuss the implication of channel kinetics for the patterning of trains of neuron spikes, a brief remark on the fractal activity at the site of neural impulse transmission at the neuromuscular junction. As is well known from the work of DelCastillo and Katz (1954), the neural transmitter substance acetylcholine is released from the nerve terminal in small packages: the miniature end potentials (MEPP) are considered manifesta tions of the exocytosis of humoral transmitters. In departure from initial textbook accounts of the MEPP release reflecting a set of homogeneous stationary Bernoulli trials, Perkel and Feldman (1979) categorically reject a purely binomial model of (quanta l) transmitter release. For the frog neuromuscular junction, Rothshenker & Rahaminoff (1970) could show that excocytosis can exhibit correlations (memory) extending over periods of seconds, suggesting self-similar characteristics. When sampled over prolonged periods, Lowen et al (1997) collected conclusive data at the neuromuscular junction and synapses in hippocampal tissue culture that frequency and amplitudes of MEPP’s display fractal scaling .Takeda et al (1999) also reported comparable findings for the vertebral neuromuscular junction. The detailed analysis of quantitative features of the recorded data led Lowen at al (1997) to conclude that traditional renewal models of vesicular exocytosis as a memoryless stochastic process are entirely inadequate for representing many of its salient features. Instead, their recommendation is that a new class of models should be considered that relies on fractal-rate stochastic point processes: fractal rate activity represents a kind of memory in that occurrence of an event at a given point in time increases the likelihood of another event to occur at a later point in time, with that likelihood persisting for some time. 2.2.2 Point process analysis in peripheral nerves and neurons In this section, neuron discharge trains are viewed as mathematical objects, belonging to the class of point processes (Thurner et al, 1997; Lowen and Teich, 2005): events occurring at a point in time or space. Werner and Mountcastle (1963, 1964) determined scaling of neural responses in primary cutaneous afferent nerve fibers with the magnitude of mechanical stimuli applied to receptors. The implications of their findings in Psychophysics will be taken up in Section 3. Adaptation in neural structures serves to extend their dynamic range: the significance of this function is discussed in Section 5.2. The statistics of action potential trains recorded from single neurons in the cochlear nucleus of anaesthetized cats formed the basis of a mathematical analysis by Gerstein and Mandelbrot (1964 ). The principal result was that a random walk model towards an absorbing and a reflecting barrier can account for a wide range of fractal neuronal activity patterns, assuming no more than the known physiological mechanisms of a threshold for membrane depolarization, and the summation of excitatory and inhibitory post synaptic potentials. Except for a thesis by Johannesma in 1969, it took almost 20 years of hegemony of Poisson and Gaussian distributions until fractal approaches to spike train statistics were resumed: this time by Wise (1981) in a study of spike interval distributions of data that had been recorded primarily by Bloom (1969) in the cerebral cortex, and in respiratory neurons recorded by Smolders and Folgering (1977. Wise found that plots of the spike interval histograms on log-log scales showed negative powers on time with long tails, which he attributed to the neuron membrane potential undergoing a random walk while the firing threshold fluctuates. Re- working some of Wise’s data, West and Deering (1994) identified fractal (hyperbolic) spike interval distributions. Taking an entirely different approach to conceptualizing irregular behavior in neuron spike trains led Shahverdian and Apkarian (1998) to discuss self -affinity, powerlaw dependence and computational complexity of spike trains in terms of a multidimensional Cantor space with zero Lebesgue measure as attractor. The turning point in the history of identifying fractal neuronal firing is associated with the work of Teich and Lowen, beginning in the early 1980s (Ch 22, in McKenna, 1992) with invalidating the then prevalent notion of Poisson point processes. More recently, the shortcoming of Poisson spike interval statistics was also pointed out by Kass and Ventura (2001) and by van Vreeswick (2001) who critizised experimental (Richmond et al, 1990) and theoretical (Ohlshausen and Field, 1998) reports for unwarrantedly assuming either Poisson neurons or rate based neurons with rate independent Gaussian noise; instead he considered a renewal model as biologically more plausible. Teich and Lowen’s essential realization was that determining long-time correlations in spike trains requires sample sizes to be appreciably larger than conventionally used. On this basis, Teich et al (1990) identified the following essential features of the time series of neural spikes recorded from cat auditory nerve fibers and the lateral superior olivary nucleus: discharge rates determined with different averaging times can exhibit self -similarity; the variance-to-mean ratio of spike number increases with sufficiently large counting time in a fractional power law fashion, with the exponent in the power law varying with the stimulus level. With these data in hand, Lowen and Teich (1993) suggested that the fractal action potential patterning in auditory nerve may be related to fractal activity in the ion channels of the sensory organs feeding into the auditory nerve: that is, the hair cells in the cochlea. This idea required to show that ion channel gating and neuronal spiking patterns are indeed causally related. Lowen et al (1999) succeeded with demonstrating this causal dependence in computational models, thus adding for the special case of the cochlear hair cells some credence to their proposal that gating patterns in sensory organ ion channels can affect discharge patterns in the sensory nerve tracts they feed. In an elegant experimental design, Teich (1977) not only ascertained a power function for the activity in retina ganglion cells and neurons in the lateral geniculate body when studied independently, but also succeeded with recording from synaptically connected pairs of retina ganglion cells and geniculate neurons. In this situation, fractal exponents for retina and target neurons in the lateral geniculate body were nearly identical. This was interpreted to mean that fractal behavior is either transmitted across synapses, or has a common origin for the synaptically connected pre- and postsynaptic structure. On the other hand, fractal activity of medullary sympathetic premotor and the synaptically connected pregangionic synmpathetic neurons is apparently generated independently (Orer et al, 2003). More support for the notion that ion channel properties play an important role for determining neuron performance comes from demonstrating a kind of memory mechanism for traces of prior activity in voltage-gated Na channels (Toib, 1998): time constants of channel recovery stand in a power function relation to duration of prior activation. The question of primary interest is of course how the dynamics of ion channels relates to the functional characteristics of a whole neuron. Gilboa et al (2005) addressed this question in a computational model of an ensemble of ion channels. In analogy to a ‘real’ neuron, this model neuron exhibits various dynamics at different time scales: a power law function recovery time scale after stimulation, temporal modulation of discharge pattern during maintained stimulation, and the dependence of adaptation to a stimulus step on the duration of the priming stimulus. The suggestive implication is that the ensemble of ion channels can exhibit properties on many scales, comparable to ‘real’ neurons, thus supporting the notion that the ‘macroscopic behavior’ of the ‘real’ neuron is, in fact, the result of cooperative fractal channel kinetics. In addition to the studies cited in foregoing paragraphs, there are numerous reports documenting fractal-rate behavior in single neuronal point processes. However, these data were generally obtained for examining spike trains for encoding stimulus properties, and they are quite heterogeneous as regards species, neural structure examined, use of anesthetics and experimental conditions. Although this imposes serious limitations on drawing inferences on general principles, I select here a few studies which applied several of the commonly agreed upon and typical indicators of fractal properties, such as self similarity of firing rate with different averaging time, increase of spike number variance-to-mean ratio with counting time, and power law scaling relating the variable of interest to the resolution of measurement. In a series of publications, Grueneis et al (1993) reported fractal properties in spike trains recorded under various conditions including REM sleep of cats. In visual cortical areas of cats and macaques, Baddeley et al (1997) observed consistently non-Poisson spike train statistics, with some displaying self-similarity. Other neural structures examined included medullary sympathetic neurons (Lewis et al, 1993) and dorsal horn of the spinal cord (Salvador and Biella, 1994). A common feature of these and other like reports not cited here, was the lack of agreement on a consistent mathematical model that would satisfactorily describe the fractal process underlying the experimental data. In a study of retina ganglion cells, Teich & Saleh (1981) suggest a shot-noise driven self exciting point process; in a later study of the same experimental object, Teich found a modulated gamma-r-renewal process satisfactory while Grueneis et al(1993 ) favor a clustering Poisson process. Mandelbrot and van Ness (1968) considered Fractal Brownian motion as candidate. Clearly, the goal of determining whether a common principle governing spike train variability could be identified, and if not then for what reason, eluded this group of investigators. Without examining specifically for manifestations of fractality, a number of investigators attempted statistical characterization of neural point processes, primarily motivated to reconcile irregularity of spike trains with their presumptive function as “code” of neural signals. But recall that Harris (2005) attributes irregularity of spike train discharges to cell assembly organization. In various modifications, the general approach chosen by Sakai et al (1999), Cateau and Reyes ( 2006), and Feng and Zhang (2001) consisted in designing model neurons to generate spike trains whose statistics would match that of “real” neurons recorded in animal experiments. Shinomoto et al (2003) recorded spike sequences from different cortical areas in awake macaques which they classified phenomenologically into different groups. Salinas & Sejnowski (2002) and Stevens and Zador (1998) assigned the principal source of discharge variability to correlations in the input feeding the examined neuron. None of these results warranted the allocation of observed or simulated spike train data to one of the probability distributions in the conventional repertoire of statistics, but Maimon and Assad (2009) at least excluded Poisson – like randomness from being a universal feature of spike time distributions in primate parietal cortex. In an exquisitely elegant experiment, Evarts (1967) followed the changes of interspike interval (ISI) histograms in premotor cortex pyramidal neurons in wakefulness, sleep and the phase of sleep associated with low-voltage fast EEG. Regrettably, his characterization of the ISI histograms is limited to rejecting Poisson distributions. However, inspecting the histograms displayed in Fig. 12 of his publication arouses one’s suspicion of a long-tail distribution for sleep activity. In a notable and very extended comparison of cortical neuron discharges in alert macaques, Shadlen & Newsome (1998) attributed to single neurons the ability to perform simple algebraic operations resembling averaging by combining inputs from several sources but they cautiously concluded that irregularity of the interspike interval distribution precludes them from reflecting information about the actual temporal structure of the synaptic input. They rejected random walk models of the kind applied by Gerstein and Mandelbrot as inadequate for capturing the statistical features of spike interval distributions, and found Poisson and various renewal processes likewise failing to yield satisfactory and consistent correspondence with recorded data. If there is one conclusion that can be drawn from the extant data on the statistics of spike interval distributions, then it is that demonstrating fractal properties in spike trains requires carefully selected conditions. Multiple convergences from incoming pathways appears to obscure characteristic statistical properties of discharges in the recipient neurons. Thus, a neuron’s intrisic connection pattern carries the burden of discharge variability. This view is reminiscent of Harris’s (2005) view that spike discharge variability may be a signature of cell assembly organization. This is perhaps also the source of futility of assigning any information bearing capacity to discharge patterns of individual neurons (see for instance: Werner, 2007a). On the other hand, the more direct a neuron’s connection pattern to peripheral sensor s is, the more distinctly are fractal discharge properties demonstrable. But the opposite also seems to be the case when neurons are embedded in a network, as the observations of El Boustani et al (2009) in Section 2.3 show. The place to look for consistent fractal properties is apparently the macroscopic, global level of brain organization (see Section 2.3). Whether and how the mesoscopic level of the next Section bridges bridges the gap is the subject of the next section 2.3 The mesoscopic level of organization Despite their relative simplicity, in vitro cultured neuronal networks are here viewed as mesocopic in the sense of representing neuron ensembles which exhibit rich spontaneous dynamical activity in the form of periodic bursting (Robinson et al, 1993; Nakanishi and Kukita, 1998; van Pelt et al, 2004; earlier references are cited in: Beggs and Plenz, 2003, 2004 ). At superficial inspection, the brief burst of activity lasting tens of milliseconds are separated by quiescence lasting up to several seconds (Corner et al, 2002; Tateno et al, 2002). The spontaneous emergence of patterns in the discharge trains was also noted by Giugliano et al (2004) and replicated in computational models . In extension of earlier work that led to identification of scale invariant Levy distributions and long range correlations in cultured neuronal networks (Segev et al, 2002; see Section 4 ), Segev et al (2004) attributed the activity bursts to be associated with the formation of statistically distinguishable subgroups of neurons , each with its own distinct pattern of interneuronal spatiotemporal correlations. Wagenaar et al (2006 a) emphasized the extremely rich repertoire of bursting patterns during the development of cortical cultures. The cultured cortical networks spontaneously generated a hierarchical structure of periodic activity with a stereotyped population -wide spatiotemporal structure. These recurring patterns ( called by the authors ‘superburst’ ) converged periodically to a dynamic attractor orbit, and were taken to imply large-scale self-organization of neurons in vitro , refuting the commonly held view of having random connectivity (Wagenaars, 2006 b). The recorded data of these authors are available for distribution to interested investigators. Departing from the practice of focusing on spontaneous activity, Breskin et al (2006) explored the propagation of stimulus evoked activity in neuron cultures. A graph theoretic analysis of their data attributed the dynamic evolution of the network connectivity to a percolation transition with power law characteristics, while the degree distribution of the grown network failed to meet power law criteria. Working with mature organotypical cultures and acute slices of rat cortex, Beggs and Plenz (2003, 2004) concluded that the cascades of propagating neuron discharges they observed were indicative of the neural culture being in a state of self-organized criticality. The importance of this claim, and a recent disputes of its validity (see below), warrant close attention to methodological details: Beggs and Plenz based their analysis on recording spontaneously appearing negative local field potentials (NLFP), apparently occurring preferentially in cortical layers 2/3 (Gireesh and Plenz, 2008). The peaks of NLFP were considered indicative of synchronized population discharges occurring near the recording electrode tip (Plenz and Aertsen, 1993), on the rational that bursts of multiunit activity are more likely to generate large NLFP’s than are single neuron discharges (Plenz and Thiagarajan, 2007). Accordingly, brief bursts of synchronized action potentials were the units of analysis in their experiments. The alternation between brief burst of NLFP activity and quiescent periods remained in their experiments stable with a high degree of temporal precision, extending over periods of many hours. Beggs and Plenz set out to examine the idea that these cascades of neural activity may constitute a special mode of network activity, possibly of the character of “avalanches”, indicative of SOC (Bak et al., 1987). They had specifically in mind the kind of self-organizing branching process discussed by Zapperi et al, ( 1995) and de Carvalho and Prado (2000). To examine this idea required determining the statistical properties of the observed activity patterns. For carrying out this analysis, they first defined the spatial pattern of signal- carrying electrodes during one time bin as frame; a sequence of consecutively active frames, preceded and ended by a blank period was called an avalanche. The NLFP did not propagate in the network in a spatially contiguous manner, thus excluding wave-like propagation. With these definitions and precautions in place, distributions of avalanche size and lifetimes were found to scale in cultures and acute cortex slices with a power law exponent -3/2, this value being resilient to various choices of scales (Plenz and Thiagarajan, 2007). The branching parameter was determined as sigma =1.04 . Being statistically indistinguishable from the ideal value of 1 , it signifies a critical state in the sense that activity starting at one electrode would initiate activity in one other electrode, on the average, keeping the network at the edge of criticality (Harris, 1989). This complements the evidence for fractal properties and supports the validity of the working hypothesis Beggs and Plenz started out with : to view avalanches as manifestations of the collective critical dynamics of SOC. More recently , fractal scaling of patterned neural activity was reported to also occur in cultivated neurons of leech ganglia and rat hippocampus (Mazzoni et al, 2007) ; and Pasquale et al, (2008) describe avalanches in dissociated neuronal cultures of cortex from embryonic rats. Avalanches were subsequently also studied in superficial layers of rat prefrontal cortex (Stewart and Plenz, 2006) and during development of cortical layer 2/3 where they display nested theta- and beta/gamma oscillations ( Gireesh and Plenz, 2008). The theory of critical states predicts (see Section 2.4.3) and experiments confirm that neuronal avalanches display a maximized dynamic range of responses to stimuli (Shew et al, 2009). Functional architecture of avalanches conformed to Small World Topology (Pajevic and Plenz, 2009). Comparing NLFP activity in vitro cortex preparations with in vivo activity of awake macaque monkey cortex, Petermann et al (2009) established that high fidelity propagation of local synchronized scale- invariant activity patterns is a robust and universal feature of cortex in awake monkeys. Furthermore, large amplitude negative field potentials (like those constituting the avalanches) spread in a cascade-like fashion through the cortical network without distortion , much like action potentials: Thiagarajan et al (2010) described these stereotypical waveforms as ‘coherence potentials’. They occur often in rapid succession as a stream of dynamical associations, suggesting the switching of the cortical network from one dynam ical state to another. The reason for viewing neuronal avalanches as manifestation of self-organized criticality was based on their fractal scaling properties for size and duration. Here just a brief reminder of the amply documented ‘family resemblance’ of fractality and SOC, to which the publications of Grinstein (1995), Chen et al (1995), Paczuki et al. ( 1996), Tebbens and Burroughs (2003) and Cessac (2004) speak, as do the model computations of Papa and da Silva (1997) , da Silva et al (1998), De Arcangelis et al (2006) and Levina et al (2007 . However, despite the ‘avalanche’ of research on mechanisms leading to scale invariance, there exist questions about the necessary conditions for establishing the self-organized critical state (Kinouchi and Prado, 1999). A non- conserving critical branching model was proposed by Juanico et al,(2007) to demonstrate that SOC can occur in mean-field sand piles, provided the branching process is coupled to a background activity of spontaneous switching between refractoriness and quiescence among system components; in the stationary state, the system can undergo a transition from a subcritical to a critical state. In an elaborate recent study, Bonachela and Munoz ( 2009) claim that non-conserving (dissipative) dynamics does not lead to bona fide criticality. Non- conserving systems are in their view not truly critical models. Instead, non-conserving systems (such as the brain) would just hover around a critical point (presumably after some form of fine- tuning) with broadly distributed fluctuations which do not disappear at the thermodynamic limit. Such systems could be fluctuating in the vicinity of the critical point, but not at it. The authors call this condition ’self-organized quasi-criticality’. Whether this is the last word in the long standing debate of conditions for criticality in SOC remains yet to be seen. Conditions for universality of 1/f scaling in dissipative self-organized criticality models were established by De Los Rios and Zhang (1999). Models predicting avalanches of neural activity include the work of Herz and Hopfield (1995) and were noted by Eurich et al (2002) in a network of globally coupled nonlinear threshold elements. Models of neural networks of non - leaky integrate-and-fire neurons exhibit over a wide range of connectivity patterns power law avalanches with an exponent closely approximating that reported by Beggs and Plenz for tissue cultures (Levina et al, 2007). In a comment to this paper, Beggs (2007) gives a lucid account of how neural networks could self-organize to operate at criticality. Critical avalanche networks can be computationally constructed by simple network growth models (Abbott and Rohrkemper, 2007). An exponent of the experimentally determined value -1.5 of avalanche size and lifetime scaling is predicted by the neural field theory of Buice and Cowan (2007). A field theoretic approach was also investigated by Freeman (2005): analysis of high resolution electroencephalograms of rabbits revealed neural fields in the form of spatial patterns in amplitude and phase modulation of gamma and beta carrier waves which distinguished positive and negative conditioning stimuli. The goal of applying field theory was in these experiments to model states and state transitions as large-scale spatial patterns of neural activity for quick changes in adjustment to different behavioral situations. The cortical states were viewed as “wave packets”, resembling frames in motion picture, stabilized in a scale free state of self-organized criticality. Recall, however, the reservations raised by Touboul and Destexhe (2009) that substantiating interdependence of fractality and self - organization requires additional evidence. In an entirely different context (namely fossil extinction), Newman (1996) shows that fractality need not be a unique indicator of SOC and criticality since an alternative simple model can account for the empirical power law relation (for an extended discussion of this, see Sections 4 and 5) Seeking to strengthen the conjecture of self-organized criticality of avalanches, Plenz and Chialvo (2010) acquired experimental evidence that the neural avalanches in superficial layers of cortex exhibit five additional quantitative aspects of their dynamics which are consistent with critical dynamics. These were: separation of time scales between triggering and the avalanching event itself; stationarity of size distribution; temporal clustering prior to and following large avalanches , resembling Omori’s law; power law decay for avalanche size in the wake of preceding large avalanches; and a fractal dimension for scaling spatial spread. The importance of these results lies first, in affirming evidence for avalanches displaying robust critical behavior; and second, in suggesting that their scale-invariant (fractal) properties do in fact reflect cortical networks being in a state of criticality. This is also supported by the observation that the exponents of simulated branching processes at near-critical branching are similar to scaling exponents characterizing oscillations in the MEG recorded alpha frequency band of humans at rest (Poli et al, 2008). Criticality in cortex will be take up in Section 2.4.3, but let me merely stress at this point that it implies the brain being poised to undergo sudden second order phase transitions to new configurations with long range correlations among its disparate constituting elements. The dynamics is universal in its independence of details at the microscopic level (Sornette, 2000; Stanley, 1987). Viewing neural activity in this framework is a fundamental departure from a wave-type oscillatory or stochastic dynamics as the more commonly considered theoretical approaches in Neuroscience. To underscore this distinction, early-stage activity in the developing retinal network can be cited as an illustrative example (Hennig et al, 2009) of the essential feature of critical dynamics: rhythmic bursts of action potential in retina ganglion cells , propagating as wave-like events across the retina surface, arise at a very specific network state which meets the criteria of the classical percolation model of statistical mechanics (Essam, 1980): the phase transition consists in separating states of purely local from global functional connectedness, the latter displays in addition conspicuous fractal properties (Stauffer and Aharony ,1991/1994). It may be revealing to contrast the failure of consistently finding fractal activity patterns in individually sampled neurons (other than those receiving relatively direct input , see Section 2.2.2) with the abundance of fractal patterns of (mesoscopic) neuron ensembles. It raises the question whether the origin of the latter may be a matter of assembly organization: note that in the records of neuron cultures, it is the concurrent activity of interconnected neurons sampled by different electrodes that forms the fractal pattern; this is of course quite different from the sampling of neurons in point process analysis, guided by chance encounters of a microelectrode with one active neuron at the time. The puzzle posed at the end of section 2.2.2 thus finds perhaps its resolution in network architecture: Teramae and Fukai (2007) describe a model that shows how the fractal property of a few individual neurons can turn into an organized communal property of an ensemble. This lesson can also be learned from models of dynamic pattern formation in neuron populations, forming fractal power spectra and power law pulse distribution (Usher and Stemmler, 1995). Similarly, network amplification of local fluctuations causes fractal firing patterns and oscillatory field potentials in two -dimensional models of leaky Integrate-and-Fire neurons ; feedback connectivity of local excitation and surround inhibition being the essential prerequisites ( Usher et al., 1994). Bedard et al. (2006 b) accept existing claims for 1/f scaling of global variables of neural activity (e.g. EEG: Freeman et al, (2003); EMG: Novikov et al., (1997); see Section 2.3) , and acknowledge them as evidence for self-organized critical states with power law distributions , much as the models of De Arcangelis et al (2006) and Levina et al (2007) suggest. They also accept the validity of claims for fractality discussed in Section 2.2.2 for various point process analyses. But they contest the legitimacy of generalizing from these disparate sources of data. The connection between 1/f frequency scaling of global variables and critical states of neural activity is, in their mind, far from firmly established. Moreover, they emphasize that the association of 1/f spectra with criticality may not be obligatory (for a review: see Giesinger, 2001; Newman, 1999). Having raised this warning flags, Bedard et al (2006) not only confirmed in their own investigations the presence of 1/f frequency scaling in EEG of cat parietal cortex (in absence of anesthesia), but showed in addition 1/f frequency scaling in bipolar records of Local Field Potentials (LFP). That bipolar LFP recordings sample relatively localized populations of neurons was shown by Destexhe et al. (1999). They turned then their attention to investigating whether this 1/f scaling reflects self-organized critical states with the result that avalanche size distributions did in their experiments not follow power law scaling, nor did interspike interval distributions of concurrently recorded single neuron activity show 1/f scaling ; rather, the distributions were consistent with a Poisson process. Accordingly, Bedard et al. reject the evidence for critical state dynamics. Instead, they proposed a model that would account for 1/f frequency scaling without being associated with critical states. Their model shows that the observed 1/f scaling can indeed be produced by a band pass filtering effect of extracellular media. This means that extracellular field potentials (such as LFP) can show power law scaling while the underlying neuronal activity per se need not be critical. El Boustani et al (2007) found in experimental data and models of cortical “activated” states also evidence for Poisson spike distributions, and absence of avalanche dynamics. The irregular states of cortical networks are thought to stem from a very high dimensional deterministic chaos. However, as alternative, it is worth recalling that Harris (2005) views spike train irregularity as one of the signatures of cell assembly organization. Investigating further the discrepancy between the Beggs-Plenz and the Bedard et al observations, Touboul and Destexhe (2009) recorded avalanches in cortex of (awake) cats, paying careful attention to the conditions of data collection: avalanche statistics of negative peaks LFP (linked to neuronal firing), positive peaks (unrelated to neuron firing) and surrogate data (obtained by random shuffling experimental data) were analyzed; avalanche criteria were those of the Beggs-Plenz studies. Essentially, time and amplitude distributions of NLFP showed power law distributions, preferably at high detection thresholds. But, positive NLFP and surrogate data (randomly shuffled peak times -- essentially equivalent to a threshold stochastic process-- can also show power-law distributed amplitude distributions. The conclusion of this study, then, was that apparent power law scaling cannot be considered as proof of self-organized criticality . At this point, a comment seems in order: the publication of Gireesh and Plenz (2007) seems to suggest that cortical layers 2/3 are the preferential site of avalanche occurrence. The data analyzed in the Touboul and Destexhe study were obtained in earlier work of Destexhe et al (1999) in parietal cortex of awake cats; there is no indicat ion that a possible role of cortical layer was considered. Whether layer specific patterns of neuronal arborization (Callaway, 2002) could be a source of the discrepancy can at this point not be determined The discrepancy between presence of 1/f scaling in ensemble neural activity (EEG and LFP) and, yet, 1/f scaling not being an intrinsic property of the individual neuron itself , that Bedard and El Boustani et al claimed to have identified, motivated El Boustani et al (2009) to adopt yet another experimental approach: to this end, they measured the scaling properties of the power spectrum of the intracellularly recorded membrane potential of individual neurons. The experiments were conducted in cat primary visual cortex in vivo , the animals being anesthetized and paralyzed. Full-field visual stimuli of varying characteristics were presented to the dominant eye to drive the cortical region under study to states with different firing characteristics. The remarkable result was that the frequency scaling of individual cells was largely determined by the visual stimulus statistics. There was no consistent relation between individual neurons’ scaling exponent and the visual stimuli, neither was there any correlation between membrane potential and the spiking scaling exponents. Various control tests and a computer model corroborate the authors’ conclusion that statistical correlations in a neuron’s input (i.e. its presynaptic activity) can modify the power law exponent of its spiking activity. Hence, it appears that modulation in a neuron’s power law exponent may reflect changes in the correlation state of the network activity. According to these findings, intrinsic cellular properties do not seem to play a major role for its scaling which reflects in the authors’ s view primarily the network context. Regarding self-organization, El Boustani and Destexhe (2009) follow the lead taken in Destexhe’s Doctoral Thesis of 1992 and observations of Korn and Faure (2003), and present new evidence in support of chaotic dynamics in EEG: sensitivity to initial conditions is of course prominent; it is also associated with broad-band power spectra and a fractal attractor dimension. The authors confront at length the puzzle that coherence and low dimensiona lity at the macroscopic level of EEG is associated with stochastic neuronal dynamics at the microscopic level. Is this comparable to conditions obtaining in thermodynamics ? Whence criticality ? In peripheral neurons, it seems to be favored by closeness to input from peripheral receptors (Section 2.2.2). At the mesoscopic level, Plenz and Chialvo’s ( 2009) analysis of avalanches in primate cortex seem to assure legitimate criticality at the mesoscopic level; yet, the work of Bedard et al (2006) and El Boustani et al (2009) raises the possibility that scaling properties of neuron activity may not be of intrinsic neuronal origin, but a consequence of network activity. The next Section will continue to ask: if and where in the nervous system, and under what conditions, does fractality and criticality in the brain originate ? 2.4 The macroscopic level of neural organization: 2.4.1: Fractals in brain networks Fractality at the macroscopic brain level should be viewed in the context of, and in reference to, the two major conceptual and observational frameworks that have come to guide neuroscience research: the network structure of cortical connectivity, and the brain’s state of criticality resulting from the complexity of nonlinear dynamic interactions among its constituents. Advances in network theory (Albert and Barabasi , 2002; Dorogovtsev, 2002; Park and Newmann, 2004 ) influenced the application of computational and graph-theoretical methods for characterizing structural brain connectivity in accord with statistical and topological criteria (Hilgetag et al 2002). Examining the columnar organization of neocortical cortex in detail, Roerig and Chen (2002) found that the number of connections to a central neuron has the shape of a long-tailed histogram, fitting a power law. On the basis of this “bio- power-law connection probability function”, Stoop and Wagner (2007) tested a range of network types for spread of synchronization among cortical columns: the superiority of the power-law connection was evident. In general, interaction among neurons and neuron ensembles by synchronization is constrained by network topology (Arenas et al, 2008), hence the relevance of network architecture for Neurodynamics. The potential role of neural synchrony for perceptual organization and conscious experience is a subject of a recent review by Uhlhaas et al (2009). There is considerable evidence that anatomical and functional connections between different cortical areas possess an intricate organization in the form of “small world networks” (Watts & Strogatz, 1998) , forming clusters of nearby cortical areas with short links, which in turn have long range connections to other clusters (Hilgetag and Kaiser, 2004; Sporns and Zwi, 2004; Sporns et al, 2004; Stam, 2004; Stam and Reijneveld, 2007). Neuroanatomical data sets permit identifying a repertoire of characteristic structural building blocks (motifs) (Sporns and Koetter, 2004). A hierarchical cluster architecture is thought to provide the structural basis for stability and diversity of functional patterns in cortical networks (Kaiser et al., 2007; Kaiser, 2008). Moreover, hierarchical modular topologies assure sustained activation in neural networks, intermediate between rapid fading and generalized activity spread. This can be considered a prerequisite for the occurrence of criticality (Kaiser and Hilgetag, 2010). Hierarchical graphs can switch between different dynamic activity patterns, depending on the level of ongoing (spontaneous) background activity (Muller-Linow et al, 2008; Hutt and Lesne, 2009). In terms of hierarchy theory, these investigations do not specifically address the implications of nested hierarchies which, however, are suggested by the finding of inter- and intra-cluster network hubs (Sporns et al, 2007). Moreover, fMRI data obtained from subjects in resting state identify strong functional connections between regions for which no direct structural connections are known (Honey et al., 2009). This finding may be an indicator of nested clustering (see Sections 2.4.3 and 6). In the absence of deliberate external stimulation, neuronal cortical dynamics displays complex spatial and temporal patterns of activity. In simulations of networks that mimick the large-scale inter-areal connection patterns of cortex, activity takes place spontaneously at multiple time scales, punctuated by episodes of inter-regional phase locking of oscillations (Honey et al, 2007). Significantly, the connections link neural populations of multiple levels of scale, from whole brain regions to local cell columns: this suggests that cortical connections may be arranged in fractal, possibly self-similar patterns. Statistical measures of a computational model of a fractal connection pattern did in fact resemble those of a real neuroanatomical data set (Sporns 2006). The computational models also show that varying fractal patterns induce strongly correlated changes in several structural and functional measures of network properties, as evidence of their interdependence. In general, scale free complex networks display self-similarity under length-scale transformations (Song et al, 2005) but not necessarily with regard to degree distribution (Kim et al, 2007), but models of scale-free networks need not necessarily be fractal. How, then, can the fractality of many naturally occurring networks come into being ? Song et al. (2006) account for the simultaneous emergence of fractality, modularity and small-world effect, as well as the scale-free property of real world networks by a multiplicative growth process: the network growth dynamics is conceived as the inverse of a renormalization procedure, whereby the network hubs accrete connections by linking with less connected nodes, which leads to a robust fractal topology. Within the small-world network clusters, functional Magnetic Imaging (fMRI) identifies a scale-free connection pattern inasmuch as the number of links per network node (the node degree) satisfies a power law relationship (Eguiluz et al, 2005). Likewise, van den Heuvel et al (2008) find In an imaging study of the resting brain, that inter-voxel connections follow power law scaling as evidence for scale free network topology, possibly associated with a small -world organization. This form of organization is associated with conserved wiring length and conducive to synchronization of activity across the network (Zhou et al,2007; see also Changizi, 2003; Section 2.1). Although citing merely a small fraction of the numerous publications concerned with relations between network topology and dynamics, this section underscores two points of relevance for the objective of this review: first, the presence of, and effect on network dynamics of hierarchic network organization (itself being of several types); and, second, effects of network fractality on network dynamics; but the functional implications of the latter, notably for criticality, require further investigation, as does the possibility of self–similar modularity in brain networks. In the case of metabolic networks, the latter is shown to affect path connections for diffusion and resistance of flows (Gallos et al., 2007). 2.4.2: Fractals and Criticality of Brain States Criticality, listed in the foregoing as the second notable feature in current thinking about global brain function designates the view that brain is under normal circumstances at the verge of undergoing a second order phase transition. This is attributed to its complex organization of a large number of components interacting via nonlinear dynamic functions. Measuring the fractal dimension of EEG records, Babloyantz (1986) related different values with differences in sleep states. With subjects acting as their own controls, inhalation anesthesia causes a noticeable increase in EEG dimensionality (Mayer-Kress and Payne, 1987). Multichannel MEG records, obtained with a SQUID show scaling with varying degrees of scale similarity , decreasing with the distance between recording channel locations (Novikov et al, 1997). Studying dynamical synchronization in the brain, Gong et al (2003) find scale invariant fluctuations of dynamical synchronization in human EEG. Linkenkaer-Hansen et al (2001) report long-range temporal correlations and scaling with 10-20 Hz brain oscillations. Pursuing this observation in more detail, Linkenkaer-Hansen et al (2003, 2004) suggest that the long-term spatial-temporal structure of the complex ongoing EEG activity may reflect a memory of the system’s dynamics extending beyond just a few seconds, possibly by a continuous modification of functional brain networks in the sense of SOC . In these tests, somatosensory stimuli attenuate temporal correlations and power law scaling behavior, suggesting that stimuli degrade the network memory of its past. The relationship to SOC was also the subject of the work of Freeman et al (2003) in measurements of temporal and spatial power spectral densities that identify EEG phenomena as fractal. Moreover, Freeman (2005) proposed a field-theoretic approach to account for scale-free neocortical dynamics. In five frequency ranges (extending from 0.5 to 48 Hz), detrended fluctuation analysis of EEG show global synchronization time series with scale free features (Stam and de Bruin, 2004); the scaling exponent differs for conditions of eye open and eye closed. Stam (2005) also reviewed the nonlinear dynamical analysis of EEG and EMG at great length. Positive and negative feedback affect the scaling exponent of EEG differentially; this was determined in a detrended fluctuation analysis (Biuatti et al., 2007). Performance in Stimulus detection of weak stimuli is best accounted for by modulation of the power law component in the power spectrum of MEG record : Shimono et al, (2007) attribute this phenomenon to the brain operating in a state of self -organized criticality which modulates the power spectral exponent to optimize responsiveness to external stimuli. Transients in EEG records can be detected as differences in fractal dimension of EEG (Arle and Simon, 1990), as can be neuropathological conditions (Paramanathan and Uthayakumar, 2008 ) , and differences in age and gender (Nikulin and Brismar, 2005). Nonlinear spectral analysis enabled Kulish et al (2006) to determine in EEG a set of generalized fractal dimensions and fractal spectra which reveal differences in subjects when replying to questions with either YES or NO. In a study of human development from infancy to 16 years of age, Thatcher et al (2009) measured phase shift duration and phase locking intervals of the EEG for computing instantaneous phase differences between pairs of electrodes; the log-log spectral plots showed 1/f distributions. The data revealed increased phase stability in local systems, paralleled by lengthened periods of unstable phase relation between distant connections. These results were taken to reflect progression towards self-organized criticality, accompanying the growth spurts from infancy to adolescence. When Listening to music Bhattacharya and Petsche (2001) find homogeneous scaling in the gamma band EEG over distributed brain areas, whereas the homogeneity is reduce at rest, or when reading text or during spatial imagination. As is well known, music has been under scrutiny for fractal properties for quite some time, see for instance : Voss, 1975; Hsu and Hsu, 1991; Boon and Decroly, 1995 ; see also below: Bianco,2007). Long-range temporal correlations in spontaneous discharge patterns of hippocampal - amygdala complex neurons show a power-law relation in epileptic patients (Bhattacharya et al, 2005): activity of individual neurons was in this study recorded by means of micro-wire electrodes that had been implanted for localization of epileptic foci , and records were taken in inter-ictal periods, with the subjects being awake . Neuronal activity in substantia nigra exhibits fractal activity in anaesthetized rats, but was strikingly absent in the dopaminergic nigrostriatal neurons with relatively constant discharge rate (Rodiguez et al, 2003). The authors consider the possibility that pathological rhythmic discharges and tremor onset may be associated with loss of the fractal pattern of nigrostriatal neurons. During paradoxical sleep and in the attentive state, neurons in the mesencephalic reticular formation of unanaesthetized cats exhibit firing patterns with 1/f spectral profile (Yamamoto et al, 1986). Kodama et al, 1988) extended this observation to discharge properties of neurons in Hippocamus and ventrobasal thalamic neurons, and suggest that 1/f structured patterns in discharge trains are indicative of spatial and temporal summation of convergence. Variations in 1/f spectra in cortical and subcortical brain structures of monkeys are apparently related to differences in emotional states (Andersen et al., 2006). In a very detailed and thorough study, Bianco et al (2007) identify the EEG time series as a (non-ergodic) renewal non-Poisson process, reflecting strong deviation from exponential decay. This startling claim is based on two premises: one, the comparison with the statistics of an entirely different physical process, namely the fluorescence intermittency in blinking quantum dots (Bianco et al, 2005); and, second, on the conjecture of the brain operating at or near a self-organized critical state. The implication is that neuron synchronization can be viewed as a kind of phase transition involving the close cooperation among many constituents of a neuron set, each individual neuron in essence losing its identity. Furthermore, the absence of exponential truncation would violate the ergodic condition (Bel and Barkai, 2005). The authors then proceed to show that compositional music belongs to the same category of processes. They finally claim that the effect of music on the human brain is in fact based on the essential identity of their respective fractal dynamics, ensuing a kind of complexity matching of the interacting brain-music systems. This aspect will be further pursued in section 5.1. Equally consequential are the inferences drawn by Allegrini et al (2008) from their EEG data. The thrust of their analysis is on measuring the time distribution of reco rded events occurring simultaneously at two or more electrodes (in their terminology: coincidences); they find that the time interval between two consecutive coincidences has a waiting time distribution corresponding to perfect 1/f noise. The theoretical analysis of this finding leads these authors to infer that the coincidences are driven by a renewal process. The electroencephalographic findings in support of 1/f scaling are supplemented by observations with brain imaging: In 1997, Zarahn et al (1997) reported BOLD time series data obtained from normal subjects at rest that exhibited a fractal power spectrum and self -similar signal contributions, with disproportionate contribution of power in the spectrum for low frequencies. The temporal variablility of brain activity in time series of fMRI data in combination with a voxel-wise analysis of scaling exponents enabled Thurner et al (2004) to distinguish different physiological states of the brain. In non-active brain regions, the voxel-profile activity is described by a random walk model; in contrast, stimulus activated brain activity is characterized as correlated fractional Browninan noise. The same group of investigators (Shimizu et al, 2004) examined fMRI time series with a multifractal method to extract local singularity (fractal) exponents: the range of Hoelder exponents in voxels with brain activation is close to 1, whereas exponents in white matter and voxels in the absence of brain activation are close to 0.5 Without further discussing at this point the far reaching implications of the non- ergodicity claim (Tsallis, 2009; Tsallis, et al, 1995), I merely alert to two publications which interpret human EEG signals in terms of a Tsallis Entropy measure (Capurro et al 1998, 1999). The common theme of studies surveyed in the following is wavelet based representations of functional magnetic imaging (fMRI) time series. Amongst others, Wornell (1993) explicated in detail the role of wavelet based representations for the power law family of processes. The remarkable feature of wavelet analysis is that it can be viewed as matching self-similar processes since the wavelet coefficients exactly reproduce, from scale to scale, the self-replicating statistical structure of such processes (Abry, 2003). Publishing with various associates since 1994, Bullmore gathered extensive experience with fractal analysis of human brain activity which led eventually to the suggesting that wavelet-based f MRI time series estimates (Bullmore et al, 2001) can be viewed as realizations of Fractional Brownian Motion , i.e. a class of fractals described by Mandelbrot & Ness (1986), characterized by zero-mean, and non-stationary and non-differentiable time functions (see Section 2). Extolling further the virtues of wavelet techniques for the purposes on hand, Bullmore et al (2004) and Maxim et al, 2004) give a meticulous account of their use of the ’discrete wavelet transform’ approach to fMRI time series evaluation. In normal subjects at rest, the time series is most parsimoniously described as Fractional Gaussian Noise, signifying a persistent long-memory fractal processes of which the Hurst Exponent is a defining parameter. Interestingly, the value of this parameter in Alzheimer subjects differs from the norm (Maxim et al, 2004). Several results from the same laboratory contribute additional facets to the notion of the active brain displaying fractal properties: Achard et al (2006, 2008) applied discrete wavelet transform analysis to fMRI time series to estimate the frequency dependence of functional connectivity between some ninety cortical and subcortical brain regions; the functional networks is dominated by a neocortical core of highly connected hubs with an exponentially truncated power law degree distribution. Dynamical analysis of brain at wavelet scales from 2 - 37 Hz show the emergence of long-range connections with execution of motor tasks (Bassett et al, 2006). Under certain conditions (e.g. age, cognitive performance, certain pharmacologic interventions) brain dynamics requires a more comprehensive description than is captured by the monofractal analysis applied in the studies cited thus far (Suckling et al., 2008. In such cases, a more comprehensive description must make allowance for scaling behavior that is governed by several local scaling exponents. Multifractal analysis can then be characterized by the histogram of the Holder. Following expenditure of cognitive effort, the brain’s fractal oscillations require several minutes for returning to baseline activity, this time depending on the task’s cognitive load; this is taken to signify the relevance of fractal scaling for adaptive task processes, in addition to the role it plays for the “resting” brain (Barnes et al, 2009). The substantial evidence for modular organization of brain networks is reviewed by Bullmore and Sporns (2009), and was subsequently further refined by Meunier et al (2009), applying a method for rapid, high-resolution modular decomposition of brain functional networks (Blodel et al, 2008). Differences between low frequency BOLD signal spectral power in task and rest periods also support the notion of fMRI reflecting meaningful brain states (Duff et al, 2008), as do the emotional task dependent fractal fluctuations in fMRI of the cerebellar vermis (Andersen et al, 2006). Brain imaging in Neuropathology has revealed significant differences between patients suffering from unawareness of ownership of one arm (Asomatognosia) and those with additional confabulations (Somatoparaphrenia): the latter patients display lesions in the medial and orbitofrontal regions, in addition to the multiple large lesions including temporo-parietal sectors which are common to both groups of patients (Feinberg et al, 2010). For the distinction of stimuli related to the self (i.e. self -referential stimuli) from those not so related, Northoff et al (2006) identify processes mediated by cortical midline structures. Alternative approaches to the question of brain criticality have a distinguished history: sudden transitions between stable states of motor behavior are well known since the pioneering observations of Haken et al. 1985; Kelso, 1995). The transitions were interpreted as manifestations of metastability in the self-organizing nonlinear dynamic system of the brain, along the theoretical lines formulated in Synergetics (Haken, 1983). Criticality in brain and behavior was first mentioned by Kelso (1984) in a brief note. Using a superconducting quantum interference device (SQUID) sensor array, Kelso et al (1992) reported a few years later their observations of spontaneous transitions in neuromagnetic field patterns which occur at a critical value of a behavioral parameter: coherent states of both brain and behavior were captured by the spatiotemporal pattern of phase relations among participating components. This was considered evidence for the brain being a pattern forming system that can switch flexibly from one coherent state to another. Chialvo credites also Varela (2001) with the vision of brain large-scale dynamical properties. Locating cortical regions associated with such phase transitions of motor behavior, Meyer-Lindenberg et al (2002) showed that TMS can induce switches between two clearly defined and distinct motor behavior patterns. Additional new evidence of aspects of critical brain behavior accrued in rapid sequence in the following years in several forms: as the scale free connection pattern of cortical networks (Eguiluz et al, 2005; Chialvo, 2004; 2008) ; as result of the avalanche analysis of Beggs and Plenz (2003,2004) at the mesoscopic level; as coordination dynamics of large scale neural circuits subserving rhythmic sensorimotor behavior (Jantzen et al, 2008); and finally from two fMRI studies at the macroscopic level, which will be the subject of the following paragraphs. Two virtually simultaneously published recent studies, using different experimental strategies, deliver seemingly firm evidence for brain criticality. Kitzbichler et al (2009) based their approach on the widely accepted view that many behavioral and cognitive states are related to coherent or phase-locked oscillations in transient neuronal assemblies (for a recent summary: Womelsdorf et al, 2007). The measures for determining phase synchronization between component processes were in their study the phase lock intervals (estimating the length of time a pair of bandpass filtered oscillations remain in phase synchronization), and the lability of global synchronization (informally analogous to the previously discussed avalanches). Applying these measures to functional MRI and MEG data recorded from normal volunteers at resting state demonstrated power law scaling of both pair wise and global synchronization. They then evaluated the performance of two models , both typically being used in nonlinear dynamics: the Ising and the Kuramoto (1984) model. Observed and model generated data were identical, provided the model system was in a critical state. Hence, the authors conclude that the brain must be in a critical state. Moreover, the critical brain dynamics obtained at frequency intervals ranging from 0.05-0.11 to 62.5-125 Hz, confirming criticality of the human brain network organization across its functional bandwidth. They consider therefore ‘Broadband Criticality’ as a characteristic property of the resting brain network functional organization Although also using the Ising model as reference point for determining brain criticality, Fraiman et al (2009) followed an entirely different approach: the issue at stake in their study was to determine whether and to what extent the dynamics of the paradigmatic two dimensional Ising model at criticality displays features that correspond to patterns encountered in the imaging of (resting) brain networks. However, unlike most prior studies of brain dynamics cited in this Section, no prior assumptions on structural connectivity of brain regions were made. Instead, network connectivity was extracted from voxel correlations: thus, networks were here defined in terms of correlations among the activity at each location (voxel in the case of the brain, and lattice site in the Ising model). Prior investigations showed that the so called ‘resting state’ (absence of overt external stimulation) is subject to a Default Network Dynamics, reflecting balanced positive and negative correlations between activity in component brain regions (Fox and Raichle, 2007; Baliki et al, 2008); this is not the case under certain abnormal conditions (Baliki et al, 2008) . The result was that the dynamics of the Ising model at criticality, as captured by the correlation networks, exhibits average statistical properties which are identical to those observed in the brain networks at resting condition. Among several other network characteristics that match critical Ising dynamics with brain dynamics was also the equality of the fraction of sites with positive and negative correlations , corroborating that the dynamics of the normally functioning brain at rest being near a critical point. In any case, the unequivocal answer to the question the investigators set out to answer was that networks derived from correlations of fMRI signals in human brains are indistinguishable from networks extracted from Ising models at critical temperature. In an important next step, Expert et al (2010) investigated the large-scale dynamical properties of resting brain by examining more closely the character of the spatio-temporal correlations: considering three successive steps in spatial coarse graining, two-point correlation functions exhibit self-similarity; self-similarity in time was revealed by 1/f frequency behavior of the power spectrum. The condition of long range correlations in space and time presupposes a dynamical system at criticality; the strong correlations across large distances are indicators of highly integrated cortical states, with nearby clusters functioning in synchrony. Apart from this principal conclusion of this study, the authors of this study also alert to a significant property of the brain networks which, as noted before, are extracted from the site- to-site temporal voxel correlations: obviously, equally oriented spins in the Ising model coalesce in large domains near the critical temperature where also nontrivial collective states emerge in the Ising model’s otherwise regular lattice. Similarly, large regions of brain activate concurrently with deactivation of other regions. How does the brain self-organize to negotiate the dynamic balance between the extreme possibilities of total quiescence and explosive massive excitation? The authors refer to a discussion of this stability problem which was already noted by Abeles (1991). It motivates their question: is it necessary to confine brain activation to structural connections linking brain regions, as is customary in most current research ? (e.g.: Hagmann et al, 2008). Take the Ising model as example: there, a change of temperature can lead to the emergence of functional collectives, in the absence of preexisting structural connections. This leads Fraiman et al. to ask: might the brain, likewise, have this capacity, as basis of a kind of adaptive coordination dynamics of the kind envisioned by Kelso and Tognoli, (2007) and Tognoli and Kelso (2009) ? Apparently, SOC, metastability and phase transitions constitute a nexus of intimately interrelated dynamic processes of which fractals and self-similarity are pivotal aspects. 2.4.3: Significance of Brain Criticality In statistical Physics, systems operating at the critical point of transition between ordered and random behavior are metastable with respect to a set of control parameters, and are capable of rapid qualitative change in response to fluctuations of external input. For systems far from equilibrium most of the analytical and numerical methods of the ‘classical’ (equilibrium) theory appear to remain valid (Sornette, 2000 ). Moreover, dissipative (open) Hamiltonian System, such as the brain, have the capacity to form “strange” attractors whose boundaries and bases have fractal properties (Aguirre et al, 2009; see Section 4) . At or near the point of phase transition, the systems exhibit complex patterns of fluctuations on all scales of space and time, as one the indicators of an impending phase transition ; another is the slowing down of relaxation processes, associated forming long range correlations for efficient functional coupling among system components: both events are anticipatory signals of impending critical transitions (Scheffer et al, 2009). Fractal clusters formed by phase transitions can be characterized in terms of correlation length (Antoniou et al, 2000) which is associated with fractal scaling of clusters of correlated elements on all scales; as a result, any intrinsic scale before phase transition is de facto ‘forgotten’ (Stinchcombe, 1989). As a corollary, the system presents at the critical transition qualitatively new properties, requ iring new macroscopic descriptors. The important feature of the organization following the phase transition is to form new objects with distinct properties. In physics, this is manifest as, for instance, the phase transition from ferro- to para-magnetism, or from water to ice (Stanley, 1999). Typically, one deals with a large collection of ‘microscopic’ constituents which, at phase transitions, arrange to a macrostate which displays qualitatively novel features and properties. The macrostate’s new properties have no referent at the microscopic level, and require new descriptors: by way of illustration, think of hardness or liquidity in the ice -water example as descriptors of new physical properties, originating de novo upon phase transition. The properties described in the foregoing are universal in the sense that the apply irrespective of the system’s constituents at the microscopic level. One of the amazing features of phase transitions is that material systems of diverse physical properties at their microscopic level form on phase transition but a small number of Universality Classes which share identical macroscopic properties ( for a discussion in relation to brain function: see Werner,2009 c). Based on a stochastic theory of neural activity, Buice and Cowan (2007) developed field theoretic methods for nonequilibrium statistical processes; their model exhibits a dynamical phase transition of the universality class of directed percolation (see Section 2.4.2). Critical Theory (Stanley, 1987; Marro and Dickman, 1999) considers reality as a hierarchy of levels, each having its own scale, its own description and a theory that accounts for that description. The scale on each level emerges from the scale on the next finer level by ignoring some of the lower level details which become invisible at the higher level scale (Laughlin, 2005 ; Sokal and Bricmont,2004)). The result is a drastic reduction of dimensionality. Coarse graining (specifically renormalization group transformation ) (Fischer, 1998) unveils self-similarity at the point of phase transition. The intimate relations between scaling, renormalization group, and long-range correlations are addressed by Perez-Mercader (2004) and Penrose (1986), the latter pointing out that the definition of fractal dimension depends primarily on the distribution of widely separated sites, telling little on sites that are close together. What is the significance of criticality ? Excitable systems at criticality exhibit an optimal dynamical range for information processing (Kinouchi and Copelli, 2006; see Shew et al, 2009, in Section 2.3) . Furthermore, a model that reproduces the typical features of systems at a critical point learns and remembers complex logical rules: learning occurs by plastic adaptation of synaptic strengths, and exhibits universal features in being independent of the specific task assigned to the system (Arcangelis and Herrmann, 2010). Finally, Phase transition in critical systems provide an universal mechanism for rapid switching between different cooperative neuron collectives. These three attributes of criticality are the reason for its rapid ly moving into center stage of current brain theory. 3. Psychological and Behavioral Processes The following overview of psychological functions with power law scaling is predicated on the notion that mental states may be viewed as macrostates emerging from EEG dynamics (Allefeld, 2009), and neurophysiological processes generally. Classical Psychophysics of Helmholtz, Fechner and Weber sought to establish dependencies of perceptual experience on properties of physical stimuli impinging on sensory organs. In 1975, Stevens reported the summary of the extensive work that led him to propose that this dependency is in many sensory modalities a power function. In neurophysiological experiments, Werner and Mountcastle (1963,1964) identified the power function scaling of responses in primary afferent cutaneous nerve fibers to mechanical indentation of peripheral receptors. Neurons of primary visual cortex (V1) exhibit a higher coding efficiency and information transmission rate for input signals with natural long term (1/f) correlations (Yu et al, 2005). Copelli et al (2002) and Kinouchi and Copelli (2006) claim that Stevens’ law (1957) for intensity of subjective sensory experiences can be attributed to dynamics in a network of excitable elements constituting the peripheral receptors, set at the edge of a phase transition, i.e.: of being in a state of crit icality. For a discussion of this view, see Chialvo (2006). Unlike dismissing the fluctuations in the performance of many psychophysical task as “noise”, Gilden ( 1997,2001) attributes them to a memory process associated with active choice and discrimination. This memory process is suggested to express itself as 1/f ‘noise’ in the three major measurement paradigms in Psychophysics: speeded judgment, accuracy of discrimination and production. The 1/f fluctuations are attributed to an intrinsic dynamics, associated with the formation of representations, comparable to the kind of memory that arises in dynamical systems as they flow forward in time, along principles outlined by Beran (1994). According to this interpretations of the psychophysical observations, cognition would generate its dynamical signature as a consequence of its own activity: this would entail a fundamental revision of what is signal and what is noise in psychophysical data. Gilden points out that” the conventional experimental design and data analysis using ANOVA does in fact bury “ one of the most important signatures of what happens when the mind is working”. Timing fluctuations in tasks requiring sensorimotor coordination display cycle -to-cycle fluctuations which, analyzed as time series, show fractal scaling of power spectra . Ding et al (2002) suggest that the reason for this lies in the multiple time scale activities of distributed neural areas that contribute to the task performance. If asked to produce random series of numbers from a given set, series with short and long range correlations are produced which in most cases exhibit a power law spectrum (Morariu et al, 2001). Van Orden et al (2003) interpret serial correlations in human cognition as evidence of self -organization. In their view, self-organization coordinates the activities of the organism across a hierarchy of time scales, producing correlated variation across time: variations in response times would then appear as a natural fractal in which larger scale deviations nest within themselves smaller (self-similar) scale deviations. Accordingly, 1/f noise is in this view not sufficient evidence for self -organized criticality, but rather its necessary consequence. Similarly, Kello et al (2007) assembled reaction time and response data which lead them to considering the 1/f scaling of their data as expression of a coordinative, metastable basis of cognitive functions. This view is in effect an extension of Van Orden’s et al (2003), shifting the genesis of 1/f scaling from self-organization to metastability: the 1/f pervasiveness in the brain would be the signature of metastability associated with cognitive functions. However, these claims are challenged by Wagenmakers et al (2005) and contrasted with the alternative that long term serial dependence in data can be explained in a number of ways , for instance by mixtures of a small number of short -range processes) (Wagenmakers et al, 2004). Applied to problem solving and insight, reasoning was viewed by Stephen and Dixon (2009) as the self-organization of novel structures: taking a particular problem solving task as example, the authors suggest that the problem solution can be viewed as a phase transition in a self-organizing system whose dynamics would be reflected in power law behavior. Implications for social psychology are reviewed by Correll (2008): cognitive effort to avoid bias in judgments reduces the scaling exponents of response times relative to less challenging tasks. Grigolini et al (2009) interpret Correll’s data to suggest that increasing the difficulty of cognitive tasks would accelerate the transition from observed 1/f noise to white noise in decision making time series. The temporal structure of many human-initiated activities can display a striking regularity. Barabasi (2005) showed that a decision-based queuing process can account for the dynamics of some human patterns of activity: when individuals execute tasks based on some perceived priority, the timing of the tasks will indicate the signatures of fractal dynamics: heavy-tailed distributions with initial fast bursts. If patterns of expression in spoken language reflect in some way the organization of brain processes, then Zipf’s law is of course the notable landmark that presages more recent fascinating reports of fractal patterns and scale-invariant word transition probabilities in spoken and written texts (Costa and Sigman, 2009; Altmann et al, 2009; Alvarex-Lacalle et al, 2007), and their extension to music (Zanette, 2008). On the basis of EMG data, it appears that some common features of patterning in language, music and syntax (Patel, 2003) can be attributed to neural activity in Broca’s area and its right hemisphere homologue (Maess et al, 2001). 3.1 Symbol processing and fractals The classical book “Language of Thought” (Fodor, 1975 ) epitomizes the framework of computation-representation of the Computational Theory of Mind. However, with adopting a dynamical perspective, it became appropriate to view ‘representation’ in terms of regions of state space, and ‘computational rules’ as attractors (Elman, 1995); the dynamics is supplied by Recurrent Neural Networks (RNN). Systems of this kind learn to recognize and generate languages after being trained on suitable examples. Surprisingly, it turned out that the induction of this ability occurs when small network parameter adjustments bring about a phase transition in the neural network’s state space. Once in a certain state, machine states for correct recognitions scale with an exponent of 1.4 (Pollack, 1991). Considering the RNN as a dynamical system, it appears that its trajectories can locate regions in phase space which support fractal dynamics: putting it in a graphical way, the phase space would seem ‘peppered’ with regions for fractal dynamics (i.e. attractors), which can be reached by the trajectories of the complex system’s dynamics. Numerous additional sources point to a close, though often not readily transparent relation between the dynamics of RNN and IFS: the principle is consistent with the observations of Pollack (1991) inasmuch as fractal sets provide a method for organizing recursive computation in a bounded state space (Tabor, 2000). Furthermore, context-free grammar computation by connectionist networks using fractal sets can generate spatial representations of symbolic sequences (Tino, 1999; Jeffrey ,1990) via IFS (Barnsley & Demko, 1985; Barnsley et al, 1989). A class of associative reinforcement learning algorithms was constructed by Bressloff and Stark (1992) as an extension of non-associative schemes in stochastic automata theory; within the IFS framework, it suggested a possibly fractal nature of the learning process. Tsuda and Kuroda (2004) recently elaborated this idea and developed a mathematical model of Cantor Coding for the formation of episodic memory in the hippocampus. The intent is here merely to draw attention to a large segment of literature, o f which the foregoing citations are but a small sample that implicates interrelations between state space dynamics of RNN and IFS in the processing of symbolic information. For clarification of this relationship, Kolen (1993, 1994) proposed that the RNN’s state dynamics itself is an IFS, as a paradigmatic case of the synergism of fractal and complex system dynamics. Levy & Pollack (2001) obtained supportive evidence in that every point in the hidden layer of the RNN is either itself part of the fractal attractor of the IFS, or has an orbit that “ends on” the attractor in a finite number of steps. Two tantalizing questions arise: one, wherein does the ‘computational’ power of a Fractal System lie? How does the self-similar structures of fractals unpack layers of ‘information” for guiding actions across many scales still eludes our comprehension. And, second, what exactly is the nature of that apparent synergism between complex system and fractal dynamics ? (see Section 5.2). 3.2 Motor Behavior and Allometric Control processes During quiet standing, the human body sways in a seemingly erratic fashion. Collins and De Luca (1994) determined that the pattern of this postural sway is exhibits intrinsic correlations which can be modeled as a system of bounded correlated random walks. This result suggested to the authors that the postural control system incorporates both open and closed loop control mechanisms. The statistics of temporal patterns in spontaneous motor activity of laboratory rodents can be replicated by the stochastic mechanism of Davidsen and Schuster (2002, see Section 4) which generates power law distributions and 1 1/f power spectrum over several decades (Anteneodo and Chialvo, 2009). The presence of long -time correlations in the stride-interval time of normal humans suggested to Hausdorff et al (1995) that the activity of walking may be a self-similar fractal. Acknowledging the complexity of locomotor activity, the authors referred to its prerequisite of coordinating inputs from motor cortex, basal ganglia and cerebellum, as well as feedback from vestibular, visual and proprioceptive sources. In the Hausdorff et al. study, the gait cycle (synonymous with stride interval) was defined as the time between consecutive heel strikes of the same foot. West & Griffin (1998) and Griffin et al (2000) took a different and novel approach to the analysis of gait patterns: using the time between consecutive maximal positive extensions of the same knee for measuring the stride interval, their data analysis was based on determining the long-time correlation properties of the stride interval time series . The Relative Dispersion (given as the ratio of standard deviation to the mean) for different levels of data aggregation captures the inter-relatedness of the data across multiple time scales. The method is described in detail by Bassingwhite et al, 1994) and is designed to answer whether the correlations are self-similar upon scaling (i.e.: identical between groups of neighbors at different time scales). The result of the data analysis was that the fluctuations of the gait cycle were self-similar with a fractal dimension of 1.25. In addition, stride-interval time series itself was in this study a random fractal, consistent with the data of Hausdorff et al. The importance of the West-Griffin results lies in showing that the correlation in their data was an inverse power law of a form similar to the allometric scaling laws found in many areas of Biology : typically, allometry establishes a relation between two properties of an organism. Historically, the idea is based on Huxley’s (1931) definition of allometric growth, describing that the different growth rates of two parts of an organism are proportional to one another. In the West-Griffin studies, the allometric principle is reflected in the constancy of the Relative Dispersion over the length of the stride interval time series which, in the present case, has the non-integer fractal dimension of 1.25. Their data raise the issue of systematic control of variability, which is generic of complex physiological processes. Unlike the familiar homeostatic control that regulates system variables by negative feedback, an allometric control system is conceived as regulating variability of a process involving multiple interactions among sensors and effectors with intricate feedback arrangements, each with its own characteristic set of frequencies and time scales. Their functions are reflected by the allometric relation which captures the process’s long term memory with power law correlations, and by the power law distributions of the system variable (West, 1999 b). The significance of this principle is documented by West (1999 a, 2006) for the numerous physiological processes which are identified as fractal, on the basis of their time series behavior. Notable examples of fractal Physiology are heart rate, bronchial air ways and body temperature variability, and integrated neural control networks. In these situations, the regulatory mechanisms constitute coupled cascades of feedback loops in systems far from equilibrium. Therapeutic interventions, commonly based on the homeostatic principle which assumes the significant system variable to be normally distributed fails to take the regulatory complexity into account and may be counterproductive. Instead, Allometric (fractional) Control based on Fractional Calculus (Podlubny, 1999) provides the appropriate approach. Applications of Fractional Calculus to modeling the interdependence and organization of complex system, such as for instance the vestibulo-oculomotor system, are illustrated by Magin (2004). Changing walking speed, using metronomically controlled walking, or aging and pathological conditions introduce stress conditions to the neural control system which requires expanding the theoretical framework. Based on the notion of a stochas tic model of human gait dynamics (Ashkenazy et al, 2002), West and Scafetta (2003) tested the model of a neural pattern generator on the data set obtained by Hausdorff et al, 1995) which they showed to exhibit slightly multifractal fluctuations. Metronome timing breaks the long-time correlations of the natural pace and generates a large fractal variability of the gait regime. The two essential features of the model required for capturing the phenomenology of the data set were that the dynamics of the system unfolds on an attractor in phase space, and that the natural frequency of the attractor is replaced by a random walk over a restricted set of frequencies which leads to the multifractal output for the dynamical model (Scafetta et al, 2009). 4. Processes that generate power law distributions Antedating the modern theory of stochastic processes, Yule (1925) proposed a model of speciation to explain the highly skewed distributions of abundances of biological genera. Thirty years later, Simon (1955) derived several related stochastic processes from relatively general probability assumptions that lead to Yule-type distributions. Their characteristic properties distinguish them from the negative binomial and Fisher’s logarithmic series. Leaving open the possibility of still other generative mechanisms for power law distributions, Simon suggests that the frequency of occurrence of this empirical distribution should not come as surprise. The preferential attachment scheme for network growth (Barabasi and Albert 1999) has stimulated the recent interest in the Yule-Simon approach in as much as Bornholdt and Ebel (2001) could show that they are closely related. The important step of introducing the notion of aging of network nodes was taken by Dorogovtsev and Mendes (2000): the probability of being linked to a newly added node is taken to be proportional to its current connectivity weighted by a power law function of its age. This motivated Cattuto et al (2006) to propose a modified Yule - Simon process that takes the full history of the system into account, applying a hyperbolic memory kernel. Simon’s conclusion that power law distributions can be derived from relatively general assumption seems to be born out by the number of mathematical models that have been proposed. A shot noise process, reviewed by Milotti (2002) is an example, as is the Reversible Markov Chain Models (Erland and Greenwood, 2007), and the Clustering Poisson Point Process (Grueneis,2001), the latter already introduced in Section 1.2. The simple stochastic mechanism of Davidsen and Schuster (2002) generates pulse trains with power law distributions of pulse intervals, and 1/f power spectra over several decades at low frequencies with an exponent close to 1. Iterated function systems (IFS) are a unified approach for generating and classifying a broad class of fractals with self-similarty (Barnsley and Demko , 1985). The Chaos Game is a generalized form of this, designating a method for generating the attractor (fixed point) of any IFS. Other Recurrence Models (Kaulakys et al, 1998, 2006) derive from a more specific frame of reference insofar as they consider random walks in complex systems that display self - organization. As alternative, Ruseckas and Kaulaskys (2010) generate 1/f noise with nonlinear stochastic differential equations. Touboul and Destexhe (2009) followed a similar route when developing their case against power law scaling of neural avalanches (see Section 2.3). Physical systems whose observable properties exhibit values which randomly exceed certain critical values are candidates for applying Extreme Value Theory: the aim of the classical form of this theory is to quantify the properties of the extremes (large or small) occurring in random sequences of independent numbers. Extremal dynamics may be applied to generate objects with fractal structure (Miller et al, 1993); as Extremal Optimization, it successively eliminates undesirable components of suboptimal problem solutions (Boettcher and Percus, 2000). The various approaches discussed in the foregoing can essentially be viewed as ad hoc (Milotti ,2002). In contrast, however, there are two types of conceptual anchors that ground power law relations explicitly in larger foundational contexts. For one of the conceptual roots, I turn to the theory of Random Walks and fractional difference equations. The continuum limit of simple random walks is diffusion and, correspondingly, expressed in the mathematics of differential equations. The simple random walk aggregates the random steps from a large number of identically distributed random variables with finite variance. However, an extensive range of investigations has made it abundantly clear that simple random walks with this statistics do not capture the richness of biological data, and for that matter other fields of investigation as well (for reviews see : West and Deehring, 1995; West 1999; Bassingthwaighte et al, 1994). A decisive step beyond simple random walks was the introduction of the concept of Continuous-Time-Random Walk (CTRW) by Montroll and Weiss (1965). Some forms of CTRW are fundamentally different from the classical diffusion model by drawing the timing of steps from waiting time distributions, or by taking steps of randomly varying length. This is for instance the case when the waiting time distribution does not possess a characteristic time scale (for instance, has a power law distribution) : in this situation, the mean square displacement and the distribution of transition rates become fractal. Processes corresponding to these and related random walk models are then referred to as fractal random walks, corresponding to anomalous diffusion which occupies an important place for studying physical processes such as transport in disordered media or non-exponential (anomalous) relaxation of, for instance, glassy media. Along these lines, Montroll and West (1979), Hughes et al (1982) and others examined a large repertoire of stochastic processes with unusual probability distributions for the displacement per step. For certain parameters, these walks have infinite spatial moments, generate fractal self-similar trajectories, have characteristic functions with nonanalytic behavior, and lead to an analog of RNG transformations. In the continuum limit, the fractal random walk leads to the Fractional Langevin Equation of motion describing trajectories, and their ensemble densities, in phase space (West, 2006). Such processes are viewed as fractional kinetics, and mathematically addressed in fractional calculus (Sokolov et al, 2002; Kleinz and Osler, 2000) and by Fractal Operators (West et al, 2003). In an application to Neuroscience, Lundstrom et al (2008) showed that neocortical pyramidal neurons’ firing rate is a fractional derivative of slowly varying stimulus parameters: neuronal fractional differentiation effectively results in adaptation with many time scales (see Section 5.2). Fractional order dynamics of brainstem vestibulo-oculomotor neurons was demonstrated by Anastasio (1994) who also suggested that simulation of fractional-order differentiators and integrators can be approximated by integer-order high- and low-pass filters, respectively. Thus, fractional dynamics may possibly be applicable to motor control systems, generally. This is also suggested by the stride-interval time series of human gait being a random fractal, indicating the role of long-time correlations in walking (West and Griffin, 1999; see Section 3.1). Mandelbrot and van Ness (1968) defined Fractional Brownian Mot ions as a family of Gaussian random functions, parametrized according to the interdependence of successive increments, with the parameter ranging from zero (Gaussian Fractional Random Walk) to infinite in Fractional Brownian motion : the latter to account for the empirical studies of random phenomena with interdependence of distant samples. The conceptual connections to scaling invariance and to the theory of renormalization(Section 2.3.3) are discussed by Quian (2003). Fractional reaction-diffusion in inhomogenous media stabilizes steady state solutions of Turing patterns (Henry and Wearne, 2000). In 1987, Shlesinger et al. introduced the Levy walk as a random walk with nonlocal memory, coupling space and time in a scaling fashion. For the alpha-stable Levy Walks, the transition probability varies with the size of the step (Montroll and West (1987). Anomalous diffusion results from a Levy Flight which is a process where the time taken to complete a transition depends on the length of the step (West et al, 1997). West et al (1994) also identified dynamical generators of Levy Statistics . In an elegant step towards unifying various classes of random walks, Zumofen and Klafter (1933) applied the framework of CTRW’s to derive Levy stable processes. The interesting properties of Levy processes include their satisfying a scaling law, self-similarity and possessing memory (Allegrini et al, 2002) . Levy (1954) also generalized the Central Limit Theorem to include those phenomena for which the second moment diverges. West and Deering (1995) and West (2006) assembled a large number of data obtained from various biological systems that satisfy Levy walk statistics. In a motor skill acquisition task, Cluff and Balasubramaniam (2009) report that probability distributions for changes of fingertip speed in pole balancing are Levy distributed. In vitro recorded spontaneous electrical activity of neuronal networks exhibits scale –invariant Levy distributions and long- range correlations (Segev et al, 2002). This is thought to enable different size networks to self- organize for adjusting their activities over many time scales. Among animal movement patterns associated with random search behavior , Levy walks outperform fractional Brownian motion (Reynolds, 2009), presumably evolved under section pressure (Bartumeus, 2007). Physical process models to account for fractal heavy-tailed distributions of traffic pattern of (information) packages in LAN’s (Local Area Networks) are based on renewal reward processes, originally applied to commodity pricing (Taquu and Levy,1986). Applied to network package traffic, the model takes into account the presence of long packet trains (“on periods”, with packages arriving at regular intervals) and long inter-train pauses (“off periods”). The superposition of many such packet trains displays on large time scales the self -similar behavior LAN’s if the “on-off” distribution has infinite variance (Willinger et al, 1995, Willinger, 2000). The second conceptual framework was already introduced in Section 2.3.3: power law distributions are among the novelties that arise in the vicinity of or at the critical point of a continuous phase transition, including criticality of the self -organized kind. This should not come as surprise since scaling reflects long-time correlations in the underlying process, analogous to the comparable re-ordering process at critical phase transitions (Wilson, 1979): both cases address a class of phenomena where events at many scales make contributions of equal importance. Significantly, the comprehensive review on Fractal structures in nonlinear dynamics by Aguirre et al. (2009) begins with the sentence ”Fractal structures appear naturally in nonlinear dynamics, in such a way that the two concepts are deeply related”. Their review draws particular attention to numerous instances in nature where attractor basin boundaries in dissipative (open) systems display fractal behavior. Giesinger’s (2001) comment is a propos: “ at one point, Bak(1996) considered SOC as universal, with scaling as consequence; it appears, however, that the balance of evidence shifted the question : why is there scale invariance in Nature ? to the question: is Nature critical ? ” For constructing theories that deal with problems that have multiple scales, the renormalization group (RNG) offers a general method. In Physics, the most frequently studied situation is ‘percolation transition’ for which Newman (2005) offers a detailed account of the origin of power law scaling: the cumulative distribution of cluster sizes forms at the critical point a power law distribution. Percolation transition is a special case under the closely interconnected family of RNG and coarse graining that entails power law distributions as a source of natural fractals (see Section 1.4). Coarse graining allows one to determine whether the phenomenon under investigation has universality, apart from scaling: Universality implies that macroscopic properties of a system are independent of the system’s particular microscopic configuration. The particular values determined for a given instantiation of the system are then not significant, apart from showing that the system scales: the theoretical foundations are extensively discussed by Essam (1980) and by Stauffer and Aharony (1991/1994). For Neuroscience, Kozma et al (2005) illustrated the potential relevance of percolation for phase transitions in models of neural populations with mixed local and global interactions, and (Werner,2009 b,c) proposed Renormalization Group Transformation as a general principle to account for functional relations between levels of neural organization. Since fractals will in both situations naturally arise, it is pertinent to ask what their ro le could be. West et al (2008) and Allegrini et al (2006) attribute to them a complexity matching function which will be the subject of review and comments in the next section 3.1. Section 5: Fractals in Action. Having reached the end of the largely phenomenological surveys of fractal scaling and associated manifestations of fractality at the conventionally distinguished levels of organization and function of nervous systems, it appears inescapable to recall the title of Barnsley’s (1993 ) book ‘Fractals everywhere”. The apparent prevalence in the nervous system is matched by the numerous manifestation in physiological systems generally (West and Deering, 1995). Is the ubiquity a sign of triviality, or the result of a generic and fundamental principle of Nature? If that is the case in Biology at least, then Nature seems to adhere to it with remarkable conservatism as, for instance, the monographs of Dewey (1999) for Molecular Biophysics and of Seuront (2009) in Ecology attest. As documented by Aguirre et al. (2009) , and is referenced in several sections of this review: the intimate relation between fractals and nonlinear dynamics in dissipative systems is apparent and well substantiated. Hence the seemingly disproportionate attention paid to phase transitions and criticality in Sections 2.2 and 2.3. However, Section 4 lists a large number of alternatives for generating fractals, some of which obviously qualifying as ‘natural’, as, for instance Levy flights. Thus, the burden of proof of attributing observed fractals to nonlinear dynamics lies on identifying the fractal boundaries or the critical phase transition that gave rise to them. This subset of natural fractal must then be viewed as manifestation of, and a consequence of, the particular dynamic regime from which they originated. 5.1: the Complexity Matching Effect (CME) The issue under consideration is the communication among complex systems generating fractal signatures. The starting point is the evidence presented in Section 1.4 that the EEG time series can be identified as a (non-ergodic) non-Poisson renewal (NPR) process, reflecting strong deviation from exponential decay. A brief account of CME will suffice at this point since a comprehensive overview of the underlying principle of CME is available in West et al, (2008). CME is concerned with the conditions under which one complex network responds to a perturbation by a second complex network: Consider a NPR network with a power law index < 2 as measure of its complexity, and apply a random signal as perturbation: this is in essence comparable to the condition of aperiodic Statistic Resonance (Gammaitoni et al, 1998). Allegrini et al (2006a,b) then generalized the conditions by applying as perturbation another complex network which also satisfies the NPR condition with power index < 2. Under these conditions, it can be shown that the effect of the perturbation is maximal if the power law indices of the interacting systems are equal. The claim is that CME, as illustrated in the foregoing, applies to a large class of NPRs such as, for instance, return times for random walks, either in regular lattices or in complex networks. 5.2: Linking actions across many scales ? Even if Nature’s conservative adherence to fractals is a valid argument in support of their functional significance, we are still in the dark as to what that function may be. Emphasizing the feature of self-similarity, we can turn to types of functions which could benefit from stacking extended ranges of space and/or time scales into one compact format; moreover, there is no privileged time scale in power law dynamics. Sensory adaptation as a change over time in the responsiveness of the sensory system to a constant stimulus is a situation of this kind (Wark et al., 2007). Adaptation with power law dependence and multiple time scales has been demonstrated in nervous systems under many different conditions. Examples come from such diverse sources as electrosensory afferent nerve fibers in weakly electric fish (Xu et al,1996), spider mechanoreceptor neurons (French and Torkkeli, 2008). The auditory sensory memory which is thought to encode stimuli on multiple time scales (Ulanovsky et al., 2004). Fairhall et al (2001, b) direct attention to the speed with which the dynamics of a neural code is optimized, even when the statistical properties of the stimuli themselves evolve dynamically over a wide range of timescales, from tens of milliseconds to minutes. The source of 1/f fluctuations in human sensorimotor coordination tasks is presumably attributable to the multiple time scale activities of neural centers (Ding et al, 2002). In visual psychophysics, adaptation to contrast follows power law dynamics (Rose and Lowe, 1982) as does the tilt aftereffect (Greenlee and Mangussen, 1987). For fractional order dynamics in adaptation , see: Lundstrom,( 2008) and Anastasio (2004) in Section 4. This fragmentary compilation of diverse observations is intended illustrate the propensity of various neural structures to respond swiftly to a range of temporal and/or spatial parameters: the image of a set of strings resonating to specific frequencies comes to mind. Could self-similar fractal structures, having stored a repertoire of responses, each specific for a particular range of temporal and/or spatial stimulus parameters, fulfil this function ? On formal grounds, Thorson and Biederman-Thorson (1974) attributed sensory adaptation to distributed relaxation processes, based on nonuniformities of local “efficacy” in the transduction process at peripheral receptors. Might this “local nonuniformity” be the expression of a self-similar functionality of receptors ? Generically, models of adaptation integrate the response of a system and feed the integrated signal back to curtail that response. The type of adaptation is determined by the properties of the integrator. Applying this principle, Drew and Abbott (2006) examined a form of power law integration with the result that the suppressive effect of repeated stimuli on successive responses are accumulated with power law decline. In distinction from other forms of integration (e.g.: exponential), power law integration has the notable feature of scale- invariance and, in their simulations, replicated the published data of Xu et al (1996), cited in the previous paragraph. On the same principle, these investigators also implemented power law adaption within a standard spiking neuron model, except for approximating the time dependence of the power-law adaptation integral for computational efficiency by a series exponentials. This principle has been successfully applied by Hausdorff and Peng ( 1996) and is the basis of Anderson’s (2001) assertion of the power law being an emergent function : it amounts to linking the noninteracting exponential processes to a cascade which reproduces the power law forgetting in power law integration. The adaptation obtained of the integrate- and-fire model with adaptation current obtained by a cascade of exponential processes matched perfectly that obtained by injecting adaptation currents to the model neurons. The implications of this successful approach are considerable: the Dell and Abbott results suggest that power law adaptation can be instantiated by a cascade of a large number of processes with ordinary exponential dynamics , covering a wide range of time constants. The cascaded model design lets the temporal stimulus dynamics set the appropriate adaptation dynamics, in virtue of the numerous exponential processes with different time scales. They present a telling argument in support of the biological significance of this mechanism of adaptation: natural stimuli vary unpredictably over a wide of time scales; instead keeping the recovery time after excitation constant, power-law adaptation allows the temporal statistics of the stimuli themselves to determine the dynamics of adaptation. The work of Toib et al ( 1998) and Gilboa et al ( 2005) referred to in Section 2.2.1 are additional examples of ‘multiscale computing’ involving fractals. Are we prepared to envision a general principle of self -organizing control structures for multiscale behavior, extracting the statistics of an unpredictable environment by way of power law integration ? The papers cited in the next paragraph speak to this question. Fusi et al (2005) applied the principle of power-law forgetting of adaptation to a cascade model for regulating the plasticity of synapses as the basis of stored memories. Note that memory strength is here represented as synaptic plasticity, not as synaptic strength, as is commonly the case. Each synapse model has two levels of synaptic strength, weak and strong. Associated with each strength is a cascade of states (in one of their models, five). The cascades introduce a range of probabilities for transition between weak and strong levels of synaptic plasticity. A complicated heuristics is built into the model design that enables combinations of states of plasticity, for instance: states with low probability are paired up with labile states, and conversely, etc. The important point is that a high level of memory storage with long retention times significantly outperforms other model designs. In a model of learning visuomotor associations that are reversed unpredictably from time to time, synaptic modification occurring on multiple time scales along the same principles can be the basis for flexible behavior; the model predictions were validated with experimental data. (Fusi et al, 2007). It appears then, that the principle of shifting between different scales on demand manifests itself in many different forms, and on different levels of neural organization. Under natural conditions, are the required exponential functions for power law integration models of self-similar structures ? 6. Summary and final thoughts The assembly of largely phenomenological data was presented to support the claim that fractal processes and properties occur at many and diverse levels of neural organizations and performance, and are functionally relevant. Several issues that must not glossed over lightly needed discussion in several places : e.g. fractality as such is not an obligatory indicator of SOC ; whether non-conservative systems may be limited to a state of ‘quasi-criticality instead of being candidates for full criticality (Section 2.2); and what the origin of natural fractals may be (Section 4). However, the close relation of fractals with critical phase transitions is beyond dispute. Despite the evidence secured by Plenz and Chialvo in specially targeted experiments, critical dynamics is still by some investigators called in question at the mesoscopic level, but the evidence for its importance and essential role at the macroscopic domain is uncontroversial and solid. Among the virtues of brain criticality discussed in Section 2.4.3., there is one that has attracted attention for the longest time: it is the ability for rapid changes of state. Phase transition in critical systems provide an universal mechanism for rapid switching between different cooperative neuron collectives. Less attention received the fact , well established in Physics, that the system presents at the critical transition qualitatively new properties, requiring new macroscopic descriptors. The important feature of the organization following the phase transition is to form new objects with distinct properties: a new Ontology. In Physics, transitions between ferro- and paramagnetism are the prototypical example. We need to ask: What, if any, is the manifestation of ontological novelty in brain phase transition ? Customarily, one tends to think in terms of integration and differentiation, brought about by the change in correlation patterns of the system components, with nearby clusters functioning in synchrony. However, this way of looking at the nature of the state change fails to meet the requirement for qualitative (ontological) novelty, as the analogy with Physics would require. In the paradigmatic cases in Physics, new aggregations are being formed on phase transition which display novel properties at a macroscopic level of description. What can we assume to happen in brain on phase transition ? The obvious answer is: partitioning into neural assemblies with fractal properties. This is of course merely analogous to the fractal patterns formed at phase transition in physical systems. But here is an essential difference between the systems studied in Physics and the brain: Take the ferro-paramagnetic phase transition again as example, it consist merely in a change of spin orientation of the elementary components. But in the case of brain, the elementary components are reactive neurons which have the potential of entering into aggregations with functional interactions. Model studies of Arcangelis and Herrman (2010) with self-organizing neuronal networks show that avalanches formed at phase transition can learn complex rules on the basis of a collective process. Under appropriate conditions, the learning dynamics is universal inasmuch as even complex rules can be acquired. Recalling Section 3.1, the relation of the state dynamics of these neural networks to Integrated Function Systems and fractal attractors would constitute a notable case of the synergism of fractal and complex system dynamics . Re-visiting the work of Beggs and Plenz (Section 2.3) and putting it bluntly: phase transition endows avalanches with novel capacities that are not analytically predictable from the original state of the system. Generalizing from this, it is evident that pursuing these and related directions in targeted studies of the aftermath of brain phase transitions is imperative for seeking to gain a full appreciation of the functional novelties it may create. Whether such studies will show a way of bridging the ultimate barrier of the epistemic cut (Pattee, 2001) that separates the domain of integrable Physics from the domain of symbolic structures remains to be seen: but this issue is, I submit, at the core of the apparent object-subject (brain-mind) duality. The limited sample of recent publication on network models cited in Section 2.4.1 is but a fraction of the large number of combinations and permutations of network design parameters that are conceivable: on the one hand, ranging from the of small -world to scale free works with various degree distributions, on other hand each type being hierarchical, modular with and without hubs, and fractal. Which one from among this collection is closest to the natural brain networks? Suggestions abound, but tantalizing questions remain still open to further inquiry. An architecture that does justice to self-similarity under renormalization appears attractive as it might unite the dynamics of network topology in one common mechanism, as I suggest in that Section. Commenting on the wealth of existing data on anatomical and functional cortical networks organization may seem like “carrying vocals to Newcastle”, notably in view of the comprehensive review prepared by Bullmore and Sporns (2009). Nevertheless, a few general considerations may be of some relevance. It is a commonplace observation that Nature loves Hierarchies: in the classical “Architecture of Complexity”, Simon(1962) has given plausible reasons for this apparent love affair. A few years later, in a discussion of the organization of complex systems (Simon, 1973), observed that the term “hierarchy” has taken a somewhat generalized meaning, divorced from its original denotation in human organizations of a vertical authority structure. He there contended that in application to complex systems , “Hierarchy” has come to denote a set of Chinese Boxes of a particular kind, usually consisting of their recursion. But then, these Chinese Box hierarchies come in variants of one form or another of ordering, e.g. in the partial ordering of trees. There is then the possibility for a variety (and sometimes ambiguous) senses in which current investigators report their finding in anatomical, fMRI and model studies when speaking of hierarchical order. The network and graph-theoretical community has developed its own criteria: modularity has been introduced a measure of a networks decomposability, for instance into community structures ( Newman and Girvan, 2004), possibly at different hierarchical levels, allowing one to “zoom in and out” for finding communities on different levels; this was discussed in detail by Meunier et al, (l.c.). Mucha et al (2009) seek to address this same objective in a new framework that encompasses coupling of individual networks via links connecting nodes of one with nodes of another network at multiple scales, thus allowin g for some nesting which escapes the Newman-Girvan method, but is accessible by the approach of Sales-Pardo et al (2007). Keeping these issues in mind is, I suggest, of importance in the definitive evaluation of many of the reports of fMRI data on cortical functional organization: while describing modularity, they may not be able to capture nesting among modules, due to methodological limitations. Yet, it is the nesting of the kind that Simon refers to as “Russian doll” ( see also the Russian Matryoshka dolls of Agnati ,l.c.), where the pressing question of self- similarity arises: It has some plausibility in view of Sporns ‘s (2006) suggestive evidence for cortical connection patterns of fractal and self-similar nature. In different contexts, self-similar community structures were noted by Guimera et al (2003) in a network of human interactions , and nesting of theta- and beta gamma oscillations was noted by Gireesh and Plenz (2008) in neuronal avalanches of developing cortical layers 2/3. The foregoing discussion was merely intended to draw attention to some of the problems involved in attempts to characterize dynamic hierarchies of a system that consists of multiple levels of organization, having dynamics within and between the entities described at each of the different levels (Lenaerts et al., 2005). Hierarchical nesting would presumambly qualify brain dynamics as a dynamic hierarchy. This is consonant with the recent findings of Honey et al (2009) of indirect functional connectivity, not accounted for by structural connections, and the variability of resting-state functional connectivity across scanning sessions and model runs. As regards Criticality, it would then be imperative to examine the properties of phase transitions in dynamically hierarchic systems. In one recent study of a kinetic Ising model, it turned out that the universality class is consistent with the two-dimensional equilibrium Ising model (Rikvold et al., 1999). Sadly, despite the resourceful and imaginative studies undertaken by many investigators during the past decade, we are still some distance away from understanding the structural and functional basis of cortical dynamics; let alone the cortical-subcortical dynamics that, according to the Global Workspace Hypothesis (Baars,1988) is now generally considered decisive for cognitive functions. Notwithstanding this disappointing state of affairs, one more speculations may perhaps be permitted: recall the amazing discovery from Section 5.2 : there, power-law forgetting of adaptation and synaptic modification on multiple time scales were discussed as principles for flexible behavior. Recall also the evidence from Section 2.2.1 for multiscale computing as basis cooperative fractal channel kinetics. The essential principle in these cases is the capacity of having available the responsiveness, on demand, to a large range of unpredictable external contingencies of a certain category, varying in one parameter: e.g. the capacity for selecting the level of adaptation appropriate to a wide range of temporal patterns and stimulus intensity. In the cases referred to, it amounted to scale shifting as required, involving fractality and self- similarity of control structures. If Sporns’s evidence for cortical fractality and self-similarity holds up, I can now re-visit the claim I made in the Introduction: “is there a natural capacity for unpacking interactions between different levels of a fractal object or process, responsive to circumstances and conditions, which eludes us ent irely? If it existed, fractals would surely be a most extraordinary design principle for operational economy in complex systems.” To which I now add: if this natural capacity embodies principles of power law integration and cascading, similar to those identified for adaptation and fractal kinetics, we may come to comprehend at least some aspects of the extraordinary functionality of the cortex. 7. References Abbott, L.F., Rohrkemper, R. (2007). A simple growth model constructs critical avalanche networks. Progress in Brain Research 165:13-19. Abeles, M. (1991) Corticonics: Neural circuits of the cerebral cortex. Cambridge University Press, Cambridge. Abry, P., (2003). Scaling and Wavelets: an introductory Walk. Lecture Notes in Physics 621:34-60. Achard, S., Salvador R., Whicher B., Suckling J., and Bullmore E. (2006). A Resilient, Low -Frequency, Small- World Human Brain Functional Network with Highly Connected Association Cortical Hubs. J Neurosci, 26(1):63- 72. Achard, S., Bassett, D. S., Meyer-Lindenberg, A., and Bullmore, E. (2008). Fractal connectivity of long -memory networks. Physical Review, E 77, 036104. Agnati, L.F., Santarossa, L., Genedani, S., Canela, E.I., Leo, G., Franco, R., Woofs, A., Lluis, C., Ferre, S., Fuxe, K. (2004) On the nested hierarchical organization of CNS: basic characteristics of neuronal molecular organization. In: Cortical Dynamics LNCS 3146, edit. P. Erdi, pp. 24-54, Springer Berlin. Agnati, L.F., Baluska, F., Barlow, P.W., Guidolin,D. (2009) Mosaic, self -similarity Logic, and Biological Attraction. Commun. Integr. Biol. 2(6):552-563. Aguirre, J., Viana, R.L., Sanjuan, M.A.F, (2009). Fractal structures in nonlinear dynamics. Reviews of Modern Physics, 81:333-386.FF Albert, R., Barabasi, A. (2002). Statistical Mechanics of complex Networks. Reviews of Modern Physics 74:47-97. Allefeld, C., Atmanspacher, H., and Wackermann, J. (2009). Mental states as macrostates emerging from brain electrical dynamics. Chaos, 19, 015102. Allegrini, P., Bellazzini, J., Bramanti, G., Ignaccolo, M., Grigolini, P., and Yang., J. (2002). Scaling breakdown: A signature of aging. Physical Review, E 66, 015101 Allegrini, P., Bologna, M., Grigolini, P., and West, B. J. (2006a). Response of Complex Systems to Complex Perturbations: the Complexity Matching Effect. arXiv:cond-mat/0612303v1 Allegrini, P., Bologna, M., Grigolini, P., and Lukovic, M. (2006b). Response of Complex Systems to Complex Perturbations: Complexity Matching. arXiv:cond -mat/0008341v1. Allegrini, P., Menicucci, D., Bedini, R., Fronzoni, L., Gemignani, A., Grigolini, P., West, B. J., and Paradisi, P. (2008). Spontaneous brain activity as a source of perfect 1/f noise. Altmann, E. G., Pierrehumbert, J. B., and Motter, A. E. (2009). Beyond word frequency: Bursts, lulls, and scaling in the temporal distributions of words. arXiv:0901.2349v1 Alvarez-Lacalle, E., Dorow, B., Eckmann, J.-P., and Moses, E. (2006). Hierarchical structures induce long -range dynamical correlations in written texts. Proc.Natl.Acad.Sci USA, 103(21):7956-7961. Anastasio, T.J. (1994). The fractional-order dynamics of brainstem vestiblo -oculomotor neurons. Biol. Cybern., 72:69-79. Andersen, C.M. (2000). From Molecules to Mind: how vertically convergent fractal time fluctuations unify cognition and emotion. Consciousness & Emotion 1(2):193-226 Andersen, C.M., Lowen, S.B., and Renshaw, P.F. (2006). Emotional task -dependent low-frequency fluctuations and methylphenidate: Wavelet scaling analysis of 1/f-type fluctuations in fMRI of the cerebellar vermis. J. Neurosci. Methods, 151:52-61. Anderson, R.B. (2001). The power law as an emergent property. Memory & Cognition 29(7): 1061-1068. Anteneodo, C., Chialvo, D.R. (2009) Unraveling the fluctuations of animal motor activity. Chaos 19:1 Antoniou, N.G., Contoyiannis, Y. F., and Diakonos, F.K. (2000). Fractal geometry of critical systems. Physical Review E, 62, 3. Arenas, A., Diaz-Guilera, A., Kurths, J., Moreno, Y., and Zhou, C. (2008). Synchronization in complex networks. Physics Reports, 469:93-153. Arle, J. E., and Simon, R. H. (1990). An application of fractal dimension to the detection of transients in the electroencephalogram. Electroencephalography and clinical Neurophysiology , 75:296-305. Ashkenazy, Y., Hausdorff, J.M., Ivanov, P.Ch., Stanley, E. (2002). A stochastic model of human gait dynamics. Physica A 316:662-670. Baars, B.J., (1988). A Cognitive Theory of Consciousness. Cambridge University Press. Babloyantz, A. (1986). Evidence of chaotic dynamics of brain activity during the sleep cycle. In: Dimensions and Entropies in Chaotic Systems - quantification of complex behavior. ed. G. Mayer-Kress, Springer. pp. 114-122 Baddeley, R., Abbott, L.F., Booth, M. C. A., Sengpiel, F., Freeman, T., Wakeman, E.A., and Rolls, E. T. (1997). Responses of neurons in primary and inferior temporal visual cortices to natural scenes. Proc. R. Soc. Lond. B, 264:1775-1783. Bak,P. (1996). How Nature works: the Science of self-organized criticality. Springer, New York Bak, P., Tang, C., and Wiesenfeld, K. (1987). Self-Organized Criticality: An Explanation of 1/f Noise. Physical Review Letters, 59(4):381-384. Bak, P., and Paczuski, M. (1995). Complexity, contingency, and criticality. Proc. Natl. Acad. Sci. USA, 92:6689- 6696. Baliki, M.N., Geha, P.Y., Apkarian, A.V., Chialvo, D.R. (2008). Beyond feeling: chronic pain hurts the brain, disrupting the default-mode network dynamics. J. Neuroscience 28(6):1398-1403. Barabasi, A-L. (2005), The Origin of bursts and heavy tails in human dynamics. Nature 435:207-211. Barnes, A., Bullmore, E.T., and Suckling, J. (2009). Endogenous Human Brain Dynamics Recover Slowly Following Cognitive Effort. PLoS ONE, 4(8), e6626. Barnsley, M.F. (1993). Fractals everywhere. Academic Press, N.Y. Barnsley, M.F., and Demko, S. (1985). Iterated function systems and the global construction of fractals. Proc. R. Soc. Lond. A, 399:243-275. Barnsley, M.F., Elton, J.H., Hardin, D.P. (1989). Recurrent iterated function systems. Constr.Approx. 5:3-31. Bartumeus, F. (2007) Levy processes in animal movement: an evolutionary hypothesis. Fractals 15(2): 151-162. Bassett, D.S., Meyer-Lindenberg, A., Achard, S., Duke, T., and Bullmore, E. (2006). Adaptive reconfiguration of fractal small-world human brain functional networks. PNAS, 103(51):19518-19523. BassingwhiteJ.B., Liebovitch, L.S., West, B.J. (1994). Fractal Physiology, Oxford University Press, New York Bedard, C., Kroeger, H., Destexhe, A. (2006a). Model of low-pass filter of local field potentials in brain tissue. Physical Review E 73: 051911 Bedard,C., Kroeger H., Destexhe, A. (2006b). Does the 1/f frequency scaling of brain signals reflect self-organized critical states? Physical Review Letters 97:118102. Beggs, J.M. (2007) How to build a critical mind. Nature Physics 3:834-835. Beggs, J.M. (2008) The criticality hypothesis: how local cortical networks might optimize information processing. Phil. Trans. R.Soc. A 366:329-343. Beggs, J. M., and Plenz, D. (2003). Neuronal Avalanches in Neocortical Circuits. J Neurosci, 23(35):11167-11177. Beggs, J. M., and Plenz, D. (2004). Neuronal Avalanches Are Diverse and Precise Activity Patterns That Are Stable for Many Hours in Cortical Slice Cultures. J Neurosci, 24(22):5216-5229. Bel, G., and Barkai, E. (2005). Weak Ergodicity Breaking in the Continuous -Time Random Walk. Physical Review Letters, 94, 240602. Beran, J. (1994). Statistics for Long-Memory Processes. Chapman & Hall. Bernard, F., Bossu, J-L., and Gaillard, S. (2001). Identification of living oligodendrocyte developmental stages by fractal analysis of cell morphology. J. Neurosci. Res., 65(5):439-445. Bhattacharya, J., and Petsche, H. (2001). Universality in the brain while listening to music. Proc. R. Soc. Lond. B, 268:2423-2433. Bhattacharya, J., Edwards, J., Mamelak, A.N., Schuman, E.M. (2005). Long -range temporal correlations in the spontaneous spiking of neurons in the Hippocampal-Amygdala complex of humans. Neuroscience 131:547-555. Bieberich, E. (2002). Recurrent fractal neural networks: a strategy for the exchange of local and global information processing in the brain. BioSystems, 66:145-164. Bianco, S., Grigolini, P., and Paradisi, P. (2005). Fluorescence intermittency in blinking quantum dots: Renewal or slow modulation? J. Chem. Physics, 123, 174704. Bianco, S., Ignaccolo, M., Rider, M. S., Ross, M. J., Winsor, P., and Grigolini, P. (2007). Brain, music, and non - Poisson renewal process. Physical Review E, 75, 061911. Blondel, V.D., Guillame, J-L, Lambiotte, R., Lefebvre, E. (2008) Fast unfold ing of communities in large networks. J. Stat.Mech.Theory E 1P10008 Bloom, J. L., (1969). Quantitative Aspects of neural activity in the cerebral cortex of the rabbit. Thesis, Amsterdam University, Netherlands. Boettcher, S., Percus, A. (2000) Nature‟s way of optimizing. Artificial Intelligence 119:275-286. Bonachella, J.A., Munoz, M.A. (2009) Self-organization without conservation: true or just apparent scale invariance ? J.Stat.Mech. 1-37. Boon, J. P., and Decroly, O. (1995). Dynamical systems theory for music dynamics. Chaos, 5(3):501-508. Bornholdt, S., and Ebel, H. (2001). World Wide Web scaling exponent from Simon‟s 1955 model. Physical Review E, 64, 0351004. Breskin, I., Soriano, J., Moses, E., and Tlusty, T. (2006). Percolation in Living Neural Networks. Physical Review Letters, 97, 188102. Bressloff, P.C., Stark J. (1992). Analysis of associative reinforcement learning in neural netwo rks using iterated function systems. IEEE Trans.Man,Cybern. 22:1348-1360. Buiatti, M., Papo, D., Baudonniere, P.-M., and Vreeswijk, C. V. (2007). Feedback Modulates the Temporal Scale - Free Dynamics of Brain Electrical Activity in a Hypothesis Testing Task. Neuroscience, 146:1400-1412. Buice, M.A., Cowan, J.D. (2007). Field-theoretic approach to fluctuation effects in neural networks. Physical Review E 75:051919. Bullmore, E., Long, C., Suckling, J., Fadili, J., Calvert, G., Zelaya, F., Carpenter, T. A., and Brammer, M. (2001). Colored Noise and Computational Inference in Neurophysiological (fMRI) Time Series Analysis: Resampling Methods in Time and Wavelet Domains. Human Brain Mapping, 12:61-78. Bullmore, E., Fadili, J., Maxim, V., Sendur, L., Whitcher, B., Suckling, J., Brammer, M., and Breakspear, M. (2004). Wavelets and functional magnetic resonance imaging of the human brain. NeuroImage, 23:S234-S249. Callaway, E.M. (2002). Cell type specificity of local cortical connections. J. Neurocytology 31:231 -237. Capurro, A., Diambra, L., Lorenzo, D., Macadar, O., Martin, M. T., Mostaccio, C., Plastino, A., Rofman, E., Torres, M. E., and Velluti, J. (1998). Tsallis entropy and cortical dynamics: the analysis of EEG signals. Physica A, 257:149-155. Capurro, A., Diambra, L., Lorenzo, D., Macadar, O., Martin, M. T., Mostaccio, C., Plastino, A., Perez, J., Rofman, E., Torres, M. E., and Velluti, J. (1999). Human brain dynamics: the analysis of EEG signals with Tsallis information measure. Physica A, 265:235-254. Caserta, F., Stanley, H. E., Eldred, W. D., Daccord, G., Hausman, R. E., and Nittman, J. (1990). Physical Mechanisms Underlying Neurite Outgrowth: A Quantitative Analysis of Neuronal Shape. Physical Review Letters, 64(1):95-98. Caserta, F., Eldred, W. D., Fernandez, E., Hausman, R. E., Stanford, L. R., Bulderev, S. V., Schwarzer, S., and Stanley, H. E. (1995). Determination of fractal dimension of physiologically chara cterized neurons in two and three dimensions. J. Neurosci. Methods, 56:133-144. Cateau, H., and Reyes, A. D. (2006). Relation between Single Neuron and Population Spiking Statistics and Effects on Network Activity. Physical Review Letters, 96, 058101. Catutto, C., Loreto, V., and Servedio, V. D. P. (2006). A Yule -Simon process with memory. arXiv:cond- mat/0608672v1. Cessac, B. (2004). Some fractal aspects of Self-Organized Criticality. arXiv:nlin/0409004v1. Changizi, M. A. (2001 a). Principles underlying mammalian neocortical scaling. Biol. Cybern., 84:207-215. Changizi, M.A. (2001 b) Universal scaling laws for hierarchical complexity in languages, organisms, behaviors and othjer combinatorial systems. J. Theoret.Biol. 211:277-295. Changizi, M.A. (2003) The brain from 25,000 feet. Kluver Publisher, Dordrecht. Chen, D., Wu, S., Guo, A., Yang, ZR. (1995). Self-organized criticality in a cellular automaton model of pulse - coupled integrate-and-fire neurons. J. Physics A 28:5177-5182. Chen, Y., Ding, M., and Kelso, J. A. S. (1997). Long Memory Processes (1/fα Type) in Human Coordination. Physical Review Letters, 79(22):4501-4504. Chialvo, D. R. (2004). Critical brain networks. Physica A, 340:756-765. Chialvo, D. R. (2006). Are our senses critical? Nature Physics, 2:301-302. Chialvo, D.R., Balenzuela, P., Fraiman, D. (2008). The brain: what is critical about it? Conf.Proc. American Institute of Physics 1028:28-45. Clauset, A., Shalizi, C. R., and Newman, M. E. J. (2009). Power -Law Distributions In Empirical Data. arXiv:0706.1062v2. Cluff, T., and Balasubramaniam, R. (2009). Motor Learning Characterized by Changing Levy Distributions. PLoS ONE, 4(6), e5998. Collins, J.J., De Luca, C.J. (1994) Random walking during quiet standing. Physical Review Letters 73(5):764-767. Copelli, M., Roque, A. C., Oliviera, R. F., and Kinouchi, O. (2002). Physics of psychophysics: Stevens and Weber - Fechner laws and transfer functions of excitable media. Physical Review E, 65, 060901. Corner, M.A., van Pelt, J., Wolters P.S., Baker, R.E.,Nuytinck, R.H. (2002). Physiological effects of sustained blockade of excitatory tynsotic transmission on spontaneously active developing neuronal networks - an inquiry into the reciprocal linkage between intrinsic biorhythms and neuroplasticity in early ontogeny. Neurosci. Biobehav. Rev. 26:127-185. Correll, J. (2008). 1/f Noise and Effort on Implicit Measures of Bias. J of Personality and Social Psych, 94(1):48-59. Cosenza, M. G., and Kapral, R. (1992). Coupled maps on fractal lattices. Physical Review A., 46(4):1850-1858. Costa, M. E., Sigman, M., and Bonomo, F. (2009). Scale -invariant transition probabilities in free word association trajectories. Da Silva, L., Papa, A. R. R., and de Souza, A. M. C. (1998). Criticality in a simple model for brain functioni ng. Physics Letters A, 242:343-348. Davidsen, L., Schuster, H.G. (2002). Simple model for 1/f noise. Physical Review E., 65:026120. De Arcangelis, L., Perrone-Capano, C., Herrmann, H.J. (2006). Self-organized criticality for brain plasticity. Physical Review Letters 96:028107, De Arcangelis, L., Herrmann, H.J. (2010). Learning as a phenomenon occurring in a critical state. arXiv:1003.1200v1 [q-biol.NC] de Carvalho, J.X., Prado C.P.C. (2000) Self-organized criticality in the Olami-Feder-Christensen model. Physical Review Letters 84:4006-4009. DelCastillo, J., Katz, B. (1954). Quantal components of the end plate potential. J. Physiol. (London) 124:560-573. Delignieres, D., Ramdani, S., Lemoine, L., Torre, K., Fortes, M., and Ninot, G. (2006). Fractal analyses for „short‟ time series: A re-assessment of classical methods. J of Mathematical Psych, 50:525-544. De Los Rios, P., Zhang, Yi-Chen. (1999). Universal 1/f noise from dissipative self-organized criticality models. Physical Review Letters 82 (3):472-475. Destexhe, A., Contreras D., Steriade M. (1999). Spatiotemporal analysis of local field potentials and unit discharges in cat cerebral cortex natural wake and sleep states. J. Neuroscience 19 (11):4595-4608. Dewey, T.G. (1999). Fractals in Moleculal Biophysics. Oxford University Press, NY. Ding, M., Chen, Y., and Kelso, J. A. S. (2001). Statistical Analysis of Timing Errors. Brain and Cognition, 48:98- 106. Dorogovtsev,S., Mendes, J. (2003). Evolution of Networks: From Biological Nets to the Internet and WWW. Cambridge University Press, N.Y. Drew, P. J., and Abbott, L. F. (2006). Models and Properties of Power -Law Adaptation in Neural Systems. J Neurophysiol, 96:826-833. Duff, E. P., Johnston, L. A., Xiong, J., Fox, P. T., Mareels, I., and Egan, G. F. (2008). The Power of Spectral Density Analysis for Mapping Endogenous BOLD Signal Fluctuations. Human Brain Mapping, 29:778-790. Eguiluz, V. M., Chialvo, D. R., Cecchi, G. A, Baliki, M., and Apkarian, A. V. (2005). Scale-Free Brain Functional Networks. Physical Review Letters, 94, 018102 Eke, A., Herman, P., Kocsis, L., and Kozak, L. R. (2002). Fractal characterization of complexity in temporal physiological signals. Physiol. Meas., 23:R1-R38. El Boustani, S., Destexhe, A. (2009) Does brain activity tem from high -dimensional chaotic dynamics ? Evidence from the human electroencephalogram, cat cerebral cortex and artificial neuronal networks. arXiv:0904:4217v1 [nlin.CD] El Boustani, S. E., Marre, O., Behuret, S., Baudot, P., Yger, P., Bal, T., Destexhe, A., and Fregnac, Y. (2009). Network-State Modulation of Power-Law Frequency-Scaling in Visual Cortical Neurons. PLoS Computational Biology, 5(9), e1000519. Elman, J.L. (1995). Language as a dynamical system. In: Mind as Motion, Port, R.F. and Gelder,T., edits., MIT Press, Cambridge, MA. Erland, S., and Greenwood, P. E. (2007). Constructing 1/ω α noise from reversible Markov chains. Physical Reviews E, 76, 031114. Essam, J.W. (1980). Percolation Theory. Rep.Progr.Phys. 43:834-912. Eurich, C.W., Herrmann, J.M., Ernst, U.A. (2002). Finite size effects of avalanche dynamics. Physical Reviews E 66:066137 Evarts, E. V. (1967). Unit activity in Sleep and Wakefulness. In:Neurosciences, eds. GC Quarton, T. Melnechuk, F.O. Schmitt, Neurosciences Research program, pp.545-556. Expert, P., Lambiotte, R., Chialvo, D.R., Christensen, K., Jensen, H.J., Sharp, D.J., Turkheimer, F. (2010) Self - similar correlation function in brain resting state fMRI. arXiv:1003.3682v1 [q-biol.NC] Fairhall, A. L., Lewen, G. D., Bialek, W., van Steveninck, R. R. (2001a). Multiple timescales of adaptation in a neural code. Adv. Neural Inform. Proc. Syst. 13:124-130. Fairhall,A.L.,Lewen, G.D., Bialek, W., de Ruyter, van Steveninck, R.R. (2001,b) Efficiency and ambiguity in an adaptive neural code. Nature 412(6849):776-777. Feinberg, T.E., Venneri, A., Simone A.M., Fan,Y., Northoff, G. (2010). The neuroanatomy of asomatognosia and somatoparaphrenia. J. Neurol. Neursurg. Psychatry 81:276 -281. Feng, J., Zhang, P. (2001). Behavior of integrate -and-fire and Hodgkin-Huxley models with correlated inputs. Phys.Rev. E, 63: 051902. Fisher, M. E. (1998). Renormalization group theory: Its basis and formulation in statistical physics. Reviews of Modern Physics, 70:653-681. Fodor, J. (1975). Language of Thought, Harvard U.P. , Cambridge, MA. Fox,M.D., Raichle,M.E. (2007) Spontaneous fluctuations in brain activity observed with functional magnetic imaging. Nat. Rev.Neurosci. 8:701-711 Fraiman, D., Balenzuela, P., Foss, J., and Chialvo, D. R. (2009). Ising -like dynamics in large-scale functional brain networks. Physical Review E, 79, 061922.. Freeman, W.J. (2005). A field-theoretic approach to understanding scale-free neocortical dynamics. Biol.Cybern. 92:350-359. Freeman, W. J., Holmes, M. D., Burke, B. C., and Vanhatalo, S. (2003). Spatial spectra of scalp EEG and EMG from awake humans. Clinical Neurophys, 114:1053-1068. French, A. S., and Torkkeli, P. H. (2008). The Power Law of Sensory Adaptation: Simulation by a Model of Excitability in Spider Mechanoreceptor Neurons. Annals of Biomedical Engineering, 36(1):153-161. Fusi, S., Drew, P. J., and Abbott, L. F. (2005). Cascade Models of Synaptically Stored Memories. Neuron, 45:599- 611. Fusi, S., Asaad, W.F., Miller, E.K., Wang, X-J. (2007). A neural circuit model of flexible sensori-motor leraning and forgetting on multiple time scales. Neuron 54:319-333. Gallos, L.K., Song, C., Havlin, S., Makse, H.A. (2007), Scaling theory of transport in complex biological networks. Proc.Nat.Acad.Sci. USA 104(19):7746-7751. Gammaitoni, L., Hanggi, P., Jung, P., Marchesoni, F. (223). Stochastic resonance. Review of Modern Physics, 70(1):223-287. Gefen, Y., Mandelbrot, B. B., and Aharony, A. (1980). Critical Phenomena on Fractal Lattices. Physical Review Letters, 45(11):855-858. Gerstein, G. L., and Mandelbrot, B. (1964). Random Walk models for the Spike Activity of a Single Neuron. Biophysical Journal, 4:41-68. Chialvo, D.R. (2004). Critical Brain Networks. Physica A 340:756-765. Gilboa, G., Chen, R., and Brenner, N. (2005). History-Dependent Multiple-Time-Scale Dynamics in a Single- Neuron Model. J Neurosci, 25(28):6479-6489. Gilden, D.J. (1997) Fluctuations in the time required for elementary decisions. Psychological Science 8(4):296-301. Gilden, D. L. (2001). Cognitive Emissions of 1/f Noise. Psychological Review, 108(1):33-56. Gireesh, E.D., Plenz, D. (2008). Neuronal avalanches organize as nested theta - and beta/gamma oscillations during development of cortical layer 2/3. Proc.Nat.Acad.Sci. USA 105:7576-7581. Gisiger, T. (2001). Scale invariance in biology: coincidence or footp rint of a universal mechanism? Biol. Rev., 76:161-209. Giugliano,M., Darbon, P., Arsiero, M.,Luescher H.R., Streit, J. (2004). Single neuron discharge properties and network activity in dissociated cultures of neocortex. J. Neurophysiology 92:977-996. Gong, P., Nikolaev, A. R., Leeuwen, C. v. (2002). Scale -invariant fluctuations of the dynamical synchronization in human brain electrical activity. Neurosci Letters, 336:33-36. Goychuk, I., and Hanggi, P. (2002). Ion channel gating: A first -passage time analysis of the Kramers type. Proc.Natl.Acad.Sci. USA, 99(6):3552-3556. Griffin, L., West, DJ., West, BJ. (2000). Random stride Intervals with memory. J. Biol. Physics 26: 185-2000. Grigolini, P., Aquino,G., Bologna, M., Lukovic, M., West, B.J. (2009). A theory of 1/f noise in human cognition. Physics A 388:4192-4204. Gruneis, F., Nakao, M., Mizutani, Y., Yamamoto, M, Meesmann M, Musha, T, (1993). Furher study on 1/f fluctuations observed in central single neurons during REM sleep. Biol.Cybern. 68:193-198. Gruneis, F. (2001). 1/f Noise, Intermittency and Clustering Poisson Process. Fluctuation and Noise Letters, 1(2):R119-R130. Hagmann,P., Cammoun,L., Gigandet, X., Meuli,R., Honey,C., Van Wedeen,J., Spornsd,O. (2008). Mapping the structural core of human cerebral cortex. PLoS Biology 6(7): e159. Haken, H. (1983). Synergetics: an Introduction : nonequilibrium phase transitions and self -organization in Physics, Chemistry and Biology. Springer, New York. Haken, H., Kelso,J.A., Bunz, H. (1985) A theoretical model of phase transitions in human hand movements. Biol. Cybern. 51: 347-356. Harris, K.D. (2005) Neural signatures of cell assembly organization. Nature Revs.Neurosci 6:399-407. Harris, T.E. (1989),The theory of branching processes. Dover, N.Y. Harrison, K. H., Hof, P. R., and Wang, S. S.-H. (2002). Scaling laws in the mammalian neocortex: Does form provide clues to function? J Neurocytology, 31:289-298. Hausdorff, J.M., Peng, C.K., Ladin, Z., Wei, J.Y., Goldberger, A.L. (1995). Is walking a random event ? Evidence for long-range c0rrelations in stride interval of human gait. J.Appl.Physiol. 78:349-358. Hausdorff, J.M., Peng, C.K. (1996) Multiscaled randomness: a possible source of 1/f noise in Biology. Phys.Rev.E 54:2154-2157 Hennig, M.H., Adams, Ch., Willshaw D., Sernagor, E. (2009). Early-stage waves in the retinal network emerge close to a critical phase transition between local and global functional connectivity. J. of Neuroscience 29(4):1077- 1086. Henry, B.I., Wearne, S.L. (2000). Fractional reaction-diffusion. Physica A 276:448-455 Herz,A.V., Hopfield J.J. (1995). Earthquake cycles and neural reverberations: collective oscillations in systems with pulse coupled threshold oscillators. Physical Review Letters 75: 1222-1225. Hilgetag, C.C., Koetter, R., Stephan, K.E., Sporns, O.(2002). Computational methods for the Analysis of Brain Connectivity. In: Computational Anatomy, edit. G.A. Ascoli, Humana Press, pp.295 -335. Hilgetag, C. C., and Kaiser, M. (2004). Clustered Organization of Cortical Connectivity. Neuroinformatics, 2:353- 360. Honey, C. J., Kotter, R., Breakspear, M., and Sporns, O. (2007). Network structure of the cerebral cortex shapes functional connectivity on multiple time scales. PNAS, 104(24):10240-10245. Honey, C.J., Sporns, O., Cammoun,L., Thiran, J.P., Meuli, R., Hagmann, P. (2009) Predicting human resting state functional connectivity from structural connectivity. Proc.Nat.Acad.Sci. USA 106(6):2035-2040. Hsu, K. J., and Hsu, A. (1991). Self-similarity of the “1/f noise” called music. Proc. Natl. Acad. Sci. USA, 88:3507- 3509. Hughes, B. D., Montroll, E. W., and Shlesinger, M. F. (1982). Fractal Random Walks. J Statistical Phys, 28(1):111- 126. Huett, M-T., Lesne, A. (2009). Interplay between topology and dynamics in excitation pattersn on hierarchical graphs. NeuroInformatics 3:1-10. Huxley, J.S. (1932). Problems of relative growth. The Dial Press, New York. Iannacone, P.M., Khokha, M. (1995). Fractal Geometry in Biological Systems. CRC Press, Boca Raton. Im, K., Lee, J.M., Kim, S.H., Lyttelton,)., Kim, S.H., Evans,A.C., K im, S.L., (2008) Brain size and cortical structure. Cerebral Cortex 18:2181-2191. Jantzen, K.J., Steinberg, F.L., Kelso,J.A.S. (2008). Coordination Dynamics of large scale neural circuitry underlying rhythmic sensorimotor behavior. J. Cogn. Neurosci. 21(12):2420-2433. Jeffrey, H.J. (1990). Chaos game representation of gene structure. Nucleic Acid Res. 18(8):2163-2170. Jelinek, H. F., Elston,N., Zietsch,B. (1995). Fractal Analysis: Pitfalls and Revelations in Neuroscience. In: Fractals in Biology and Medicine, eds. G.A. Losa, D., Merlini, T.F. Nonnenmacher, E,.R. Weibel. Birkhauser, Basel. Jelinek, H. F. and Elston, G. N. (2001). Pyramidal Neurones in Macaque Visual Cortex: Interareal Phenotypic Variation of Dendritic Branching Patterns. Fractals, 9(3):287-295. Jones, C. L., and Jelinek, H. F. (2001). Wavelet Packet Fractal Analysis of Neuronal Morphology. DOI 10.1006/meth.2001.1205. Juanico, D.E., Monterola, C., Saloma, C. (2007) Self-organized critical branching in systems that violate conservation laws. New J. Physics 9:1-17 Kadanoff, L. P. (1990). Scaling and Universality in Statistical Physics. Physica A, 163:1-14. Kadanoff, L.P., Nagel, S.R., Wu, L., Zhou, S. (1989). Scaling and universality on avalanches. Physical Review A 39 (12):6524-6537. Kaiser, M., Goerner, M., Hilgetag, C.C. (2007) Criticality and spreading dynamics in hierarchical cluster networks without inhibition. New J. Physics 9:110-123. Kaiser, M. (2008). Brain architecture: a design for natural computation. arXiv:0802.4010v1 [q -biol. NC] Kaiser, M., Hilgetag,C.C. (2010). Optimal hierarchical modular topologies for producing limited sustained activation of neural networks. Front.Neuroinform. 4:8 Karperien, A. L., and Jelinek, H. F. (2008). Box-Counting Analysis of Microglia Form in Schizophrenia, Alzheimer‟s Disease, and Affective Disorder. Fractals, 16(2):103-107. Kass, R. E., and Ventura, V. (2001). A Spike -Train Probability Model. Neural Computation, 13:1713-1720. Katsaloulis, P.,Verganelakis, D.A. (2009). Fractal Dimension and Lacunarity of Tractography Images of the human Brain. Fractals 17 (2):181-189. Kaulakys, B., and Meskauskas, T. (1998). Modeling 1/f noise. Physical Review E, 58(6):7013-7019. Kaulakys, B., Ruscekas, J., Gontis, V., Alaburda, M. (2006). Nonlinear stochastic models of 1/f noise and power - law distributions. Physica A, 365:217-221. Kello, C. T., Beltz, B. C., Holden, J. G., and Van Order, G. C. (2007). The Emergent Coordination of Cognitive Function. J Experimental Psych, 136(4):551-568. Kelso, J.A.S. (1984) Phase transitions and critical behavior in human bimanual coordination. American Journal of Physiology: Regulatory, Integrative and Comparative, 15, R1000-R1004. Kelso, J.A.S., Bressler, S.L., Buchanan, S., DeGuzman, G.C., Ding, M., Fuchs, A. (1992).A phase transition in human brain and behavior. Physics Letters A 169:134-144. Kelso, J.A.S. (1995) Dynamic Patterns: The Self Organization of Brain and Behavior . Cambridge: MIT Press. Kelso, J.A.S., Tognoli, E. (2007).Toward a complementary neuroscience: metastable coordination dynamics of the brain. In: Kozma, R., Perlovsky, L.,(Eds). Neurodynamics of Cognition and Consciousness,, Springer, Perlin. Pp.39 - 59. Kim, J. S., Goh, K. I., Kahng, B., and Kim, D. (2007). Fractality and self-similarity in scale-free networks. New Journal of Physics, 9, 177. Kinouchi, O., Prado, C.P.C. (1999) Robustness of scale invariance in mpodels with self -organized criticality. Phys.Rev.E 59(5):4964-4969. Kinouchi, O., Copelli, C. (2006). Optimal dynamical range of excitable networks at criticality. Nature Physics 2:348-352. Kitzbichler, M. G., Smith, M. L., Christensen, S. R., and Bullmore, E. (2009). Broadband Criticality of Human Brain Network Synchronization. PLoS Computational Biology, 5(3), e1000314. Kleinz, M., and Osler, T. J. (2000). A Child‟s Garden of Fractional Derivatives. The College Mathematics Journal, 31:82-88. Kniffki, K.-D., Pawlak, M., and Vahle-Hinz, C. (1993). Scaling Behavior of the Dendritic Branches of Thalamic Neurons. Fractals, 1(2):171-178. Kniffki, K.-D., Pawlak, M.,Vahle-Hinz,C. (1994). Fractal Dimension and Dendritic Branching of Neurons in the soimatosensory Thalamus. In: Fractals in Biology and Medicine, edit.Nonnenmacher T.F. , Losa G., Weibel, E.R., Birkhauser Verlag, Basel, pp.221-229. Kodama,T., Mushiake,H., Shima, K., Nakahama, H., Yamamoto, M. (1989). Slow fluctuations of single unit activities of hippocampal and thalamic neurons in cats. I. Relation to natural sleep and alert states. Brain Research 487: 26-34. Kolen, J.F. (1994) Fool‟s Gold: Extracting finet State Machines from recurrent network dynamics. Adsv.Neur . Inform. Proc. Syst. 6:501-508. Korn, H., Faure, P. (2003) Is there chaos in the brain ? II. Experimental evidence and related models. C.R.Biol. 326: 787-840. Koulakov, A.A. (2010). On the scaling law for cortial magnification factor. arXiv:1002.4368v1[q -biol.NC] Koutsoyiannis, D. (2002). The Hurst phenomenon and fractional Gaussian noise made easy. Hydrological Sciences, 47(4):573-595. Kozma, R., Puljic, M., Ballister, P., Bollobas, B., Freeman, W. J. (2005).Phase transitions in the neuropercolation model of neural populations with mixed local and non-local interactions. Biol. Cybern., 92:367-379. Kuikka, J., Tiihonen, J, (1998). Fractal analysis – a new approach to receptor imaging. Annals of Medicine 30:242- 248. Kulish, V., Sourin, A., Sourina, O. (2006). Human electroencephalograms seen as fractal time series: Mathematical analysis and visualization. Computers in Biology and Medicine, 36:291-302. Kuramoto, Y. (1984). Chemical oscillations, waves and turbulence. Springer, Berlin. LaBarbera, M. (1989). Analyzing Body Size as a Factor in Ecology and Evolution. Annu. Rev. Ecol. Syst., 20:97- 117. Laughlin, R.B. (2005). A different Universe:Reinventing Physics from the bottom down. Basic Books, New York. Lee, J.-M., Hu, J., Gao, J., Crosson, B., Peck, K. K., Wierenga, C. E., McGregor, K., Zhao, Q., and White, K. D. (2008). Discriminating brain activity from task-related artifacts in functional MRI: Fractal scaling analysis simulation and application. NeuroImage, 40:197-212. Lenaerts, T., Chu, D., Watson, R. (2005). Dynamical hierarchies. Artificial Life 11:403-405. Levina,A., Herrmann, J.M., Geisel, T. (2007). Dynamical synapses causing self-organized criticality in neural networks. Nature Physics 3:857-860. Levina,A., Herrmann, J.M., Geisel, T. (2009). Phase transition towards criticality in a neural system with adaptive interactions. Physical review Letters 102: 118110. Levy, P., Theorie de l’addition des variables aleatoires. Gauthier-Villars, Paris Levy, S.D., Pollack J.B. (2002). Logical computation on a fractal neural substrate. http://www.demo.cs.btandeis.edu/papers/raam-ijcnn01.ps.gz Lewis, C. D., Gebber, G. L., Larsen, P. D., and Barman, S. M. (2001). Long -Term Correlations in the Spike Trains of Medullary Sympathetic Neurons. Downloaded from jn.physiology.org. Liebovitch, L. S., Fischbarg, J., Koniarek, J. P., Todorova, I., and Wang, M. (198 7). Fractal model of ion-channel kinetics. Biochimica et Biophysica Acta, 896:173-180. Liebovitch, L. S., Scheurle, D., Rusek, M., and Zochowski, M. (2001). Fractal Methods to Analyze Ion Channel Kinetics. Methods, 24:359-275. Linkenkaer-Hansen, K., Nikouline, V. V., Palva, J. M., and Ilmoniemi, R. J. (2001). Long -Range Temporal Correlations and Scaling Behavior in Human Brain Oscillations. J Neurosci, 21(4):1370-1377. Linkenkaer-Hansen, K. (2003). Scaling and Criticality in Large-Scale Neuronal Activity. Lecture Notes in Physics, 621:324-338. Linkenkaer-Hansen, K., Nikulin, V. V., Palva, J. M., Kaila, K., and Ilmoniemi, R. J. (2004). Stimulus -induces change in long-range temporal correlations and scaling behaviour of sensorimoto r oscillations. European Journal of Neuroscience, 19:203-211. Lowen, S.B., Teich, M.C. (1993). Fractal auditory nerve firing patterns may derive from fractal switching in sensory hair cell ion channels. AIP Conf. Proceedings 285 , eds. P.H. Handel, A.L. Chung, American Institute of Physics. Pp. 745-748. Lowen, S. B., Cash, S. S., Poo, M.-m., and Teich, M. C. (1997). Quantal Neurotransmitter Secretion Rate Exhibits Fractal Behavior. J Neurosci, 17(15):5666-5677. Lowen, S. B., Liebovitch, L. S., and White, J. A. (1999). Fractal ion-channel behavior generates fractal firing patterns in neuronal models. Physical Review E, 59(5):5970-5980. Lowen, S. B., Teich, M.C. (2005). Fractal based Point Processes. Wiley, N.Y. Lundstrom, B. N., Higgs, M.H., Spain, W.J., Fairhall, A. (2008). Fractional differentiation by neoircortical pyramidal neurons. Nature Neuroscience doi: 10.1038/nn.2212 Maess, B., Koelsch, S., Gunter, T. C., and Friederici, A. D. (2001). Musical syntax is processed in Broca‟s area: an MEG study. Nature Neuroscience, 4(5):540-545. Magnussen,S., Greenlee, M.W. (1985). Marathon adaptation to spatial contrast: saturation in sight. Vision Res. 25:1409-1411. Maimon, G., and Assad, J. A. (2009). Beyond Poisson: Increased Spike -Time Regularity across Primate Parietal Cortex. Neuron, 62:426-440. Mandelbrot, B. B., and Ness, J. W. V. (1968). Fractional Brownian Motions, Fractional Noises and Applications. SIAM Review, 10(4):422-437. Mandelbrot, B. B.(1977). The fractal Geometry of Nature. Marro, J., Dickman R. (1999). Nonequilibrium phase transitions in Lattice Models. Cambridge University Press, Cambridge, UK Maxim, V., Sendur, L., Fadili, J., Suckling, J., Gould, R., Howard, R., and Bullmore, E. (2005). Fractional Gaussian noise, functional MRI and Alzheimer‟s disease. NeuroImage, 25:141-158. Mazzoni, A., Broccard, F. D., Garcia-Perez, E., Bonifazi, P., Ruaro, M. E., and Torre, V. (2007). On the Dynamics of the Spontaneous Activity in Neuronal Networks. PLoS ONE, 5, e349. McKenna, T. M., (1992). Single Neuron Computation. Elsevier. Mayer-Kress, G., Layne, S.P. (1987). Dimensionality of the Human Electroencephalogram. Ann. N.Y. Acad. Sci. 504:62-87. Meunier, D., Lambiotte, R., Fornito, A., Ersche, K.D., Bullmote, E.T. (2009) Hierarchical modularity in human brain functional networks. Frontiers in Neuroinformatics 3: Article 37. Meyer-Lindenberg, A., Ziemann, U., Hajak, G., Cohen, L., and Berman, K. F. (2002). Transition between dynamical states of differing stability in the human brain. Proc.Natl.Acad.Sci. (USA), 99(17):10948-10953. Miller, S.L., Miller W.M., McWhorter, P.J. (1993) Extremal Dynamics: a unifying physical explanation of fractals, 1/f noise and activated processes. J.Appl.Phys. 73(6):2617-2628. Millhauser, G. L., Salpeter, E. E., and Oswald, R. E. (1988). Diffusion models of ion -channel gating and the origin of power-law distributions from single-channel recording. Proc. Natl. Acad. Sci. USA, 85:1503-1507. Milosevic, N. T., Ristanovic, D., Stankovic, J. B. (2005). Fractal analysis of laminar organization in spinal cord. J. of Neuroscience Methods, doi: 10.1016/j.neumeth.2005.02.009 . Milosevic, N. T., Ristanovic, D., Stankovic, J. B., and Gudovic, R. (2007). Fractal Analysis of Dendritic Arborisation Patterns of Stalked and Islet Neurons in Substantia Gelatinosa of Different Species. Fractals, 15(1):1- 6. Millotti, E. (2002). 1/f noise: a pedagogical review. arXiv:physics/0204033v1. Mitzenmacher, M. (2003). A Brief History of Generative Models for Power Law and Lognormal Distributions. Internet Mathematics, 1(2):226-251. Montroll, E. W., Weiss, G. (1965). J. Math. Phys. 6:178. Montroll, E.W., West, B.J. (1979): An enriched collection of stochastic processes. In Fluctuation Phenomena, eds. E.W. Montroll, J. Lebowitz, North Holland. Morariu, V. V., Cosa, A., Chis, M. A., Isvoran, A., Morariu, L. -C. (2001). Scaling in Cognition. Fractals, 9(4):379- 391. Mucha, P.J., Richardson, T., Macon, K., Porter, M.A. (2009). Community structure in time -dependent, multiscale and multiplex networks. arXiv:0911.1824v1 [physics.data -an] Mueller-Linhow,M., Hilgetag, C.C., Huett, M-T. (2008). Organnization of excitable dynamics in hierarchical biological networks. PLoS Comput.Biol. 4(9): e1000190. Muzy,J.F., Bacry, E., Arneodo, A.(1993). Multifractal formalism for fractal signals: the structure-function approach versus wavelet transform modulus maxima methods. Physical Review E 47:875-884. Nakanishi, K., Kukita, F. (1998). Functional synapses in synchronized bursting of neocortical neurons in culture. Brain Research 795:137-146. Newman, M.E.J. (1996). Self-organized criticality, evolution and the fossil extinction record. Proc.R.Soc.London B 263:1605-1610. Newman, M. E. J. (2003). The structure and function of complex networks. SIAM Review, 45:157-256. Newman, M. E. J. (2005). Power laws, Pareto distributions and Zipf‟s law. Contemporary Physics, 46(5):323-351. Newman, M.E.J., Girvan, M. (2004). Finding sand evaluating community structure in networks. Phys.Rev. E. 69:026113. Nikulin, V. V., and Brismar, T. Long-range Temporal Correlations in Electroencephalographic Oscillations: Relation to Topography, Frequency Band, Age and Gender. Neuroscience, 130:549-558. Northoff, G., Heinzel, A., de Greck, M., Bermpohl, F., Dobrowolny,H., Panksepp, J. (2006). Self -referential processing in our brain-- a meta-analysis of imaging studies on the self. NeuroImage 31:440-457, Novikov, E., Novikov, A., Shannahoff-Khalsa, D., Schwartz, B., Wright, J. (1997). Scale-similar activity in the brain. Physical Review E, 56(3), R2387. Ohlshausen, B. A., Field, D.J. (1997). Sparse Coding with an overcomplete Basic Set: a strategy employed by V1? Vision Res. 37:3311-3325. Orer, H. S., Das, M., Barman, S. M., and Gebber, G. L. (2003). Fractal Activity Generated Independently by Medullary Sympathetic Premotor and Preanglionic Sympathetic Neurons. J Neurophysiol, 90:47-54. Paczuki, M., Maslov S., Bak, P. (1996). Avalanche dynamics in evolu tion, growth and depinning models. Physical Review E 53(1):414-443. Pajevik, S., Plenz D. (2009). Efficient network reconstruction from dynamical cascades identifies small world topology of neuronal avalanches. PLoS ComputationalBiology 5(1): e1000271. Papa, A. R. R., Silva, L. d. (1997). Earthquakes in the Brain. Theory Bioscienc., 116:321-327. Paramanathan, P., Uthayakumar, R. (2008). Application of fractal theory in analysis of human electroencephalographic signals. Computers in Biology and Medicine, 38:372-378. Patel, A. D. (2003). Language, music, syntax, and the brain. Nature Neuroscience, 6(7):674-681. Pattee, H.H. (2001) The physics of symbols: bridging the epistemic cut. BioSystems 60:5-21 Park, J. P., and Newman, M. E. J. (2004). Statistical mechanics of networks. Physics Review E, 70, 066117. Pasquale, V. P., Massobrio, P., Bologna, L. L., Chiappalone, M. and Martinoia, S. (2008). Self -Organization and Neuronal Avalanches in Networks of Dissociated Cortical Neurons. Neuroscience, 153:1354-1369. Pellionisz, A.J. (1989). Neural Geometry: towards a fractal Model of Neurons. In: Models of Brain Function, edit. R.M.L. Cotterill, Cambridge University Press. Penrose, O. (1986). Phase Transitions on Fractal Lattices with Long-Range Interactions. J Statistical Physics, 45(1,2):69-88. Perez-Mercader, J.(2004). Coarse-graining, Scaling and Hierarchies. In: Nonextensive Entropy-interdisciplinary applications. Eds: M. Gell-Mann, C. Tsallis, Oxford Univ. Press. Pp. 357-376. Perkel, D. H. and Feldman, M. W. (1979). Neurotransmitter Release Statistics: Moment Estimates for Inhomogeneous Bernoulli Trials. J Math. Biology, 7:31-40. Petermann, T., Thiagarajan, T. C., Lebedev, M. A., Nicolelis, M. A. L., Chial vo, D. R., and Plenz, D. (2009). Spontaneous cortical activity in awake monkeys composed of neuronal avalanches. Proc. Natl. Acad. Sci. USA, 106(37):15921-15926. Plenz, D., Aertsen A., (1996) Neuronal dynamics in cortex-striatum cultures II. Spatio-temporal characteristics of neuronal activity. Neuroscience 70: 893-924. Plenz, D., Thiagarajan, T. C. (2007). The organizing principles of neuronal avalanches: cell assemblies in the cortex? TRENDS in Neurosciences, 30(3):101-110. Plenz,D., Chialvo,D.R. (2010). Scaling properties of neuronal avalanches are consistent with critical dynamics. arXiv:0912.5369v1 [q-bio.NC]. Podlubny, I. (1999) Fractional Differential Equations, Academic Press, San Diego Poli, S.-S., Ooyen, A. v., Linkenkaer-Hansen, K. (2008). Avalanche Dynamics of Human Brain Oscillations: Relation to Critical Branching Processes and Temporal Correlations. Human Brain Mapping, 29:770-777. Pollack, J.B. (1991) Induction of dynamical recognizers. Machine Learning 7:227-252.Press, W. H. (1978). Flicker Nosies in Astronomy and Elsewhere. Comments Astrophys., 7(4):103-119. Quian, H. (2003).Fractional Brownian motion and Fractional Gaussian Noise. Lecture Notes in Physics 621:22-33 Reynolds, A.M. (2009) Scale-free animal movement patterns: Levy walks outperform fractional Brownian motions and fractional Levy motions in random search scenarios. J. Phys. A 42:434006. Richmond, B. J., Optican, L.M., Spitzer, H.(1990). Temporal encoding of two-dimensional patterns by single units in Primary Visual Cortex I. Stimulus-Response relations. J. Neurophysiol. 64:351. Rikvold, P.A., Kornoiss, G., White, C.J., Novotny, M.A., Sides,S.W. (1999). arXiv:cond-mat/9904028v2 [cond- mat.stat.mech] Robinson, H.P.C., Kawahara, M., Jimbo, Y., Torimitsu, K., Kuroda, Y., Kawana, A. (1993). Periodic synchronized bursting and intracellular calcium transients elicited by low magnesium in cultured cortical neurons. J. Neurophysiology 70 (4):1606-1616. Rodriguez,M., Pereda, E., Gonzalez, J., Abdala,P., Obeso J. A.(2003). Neuronal activity in the Substantia Nigra in the anesthetized rat has fractal characteristics: evidence for firing code patterns in the basal ganglia. Exp. Brain. Res. 151:167-172. Roerig, B., Chan, B. (2002). Relationships of local inhibitory and excitatory circuits to orientation preference maps in Ferret visual Cortex. Cerebral Cortex 12:187-198. Roncaglia, R., Mannella, R., and Grigolini, P. (1994). Fractal Properties of Ion Channels and Diffusion. Mathematical Biosciences, 123:77-101. Rose, D., Lowe, I. (1982). Dynamics of adaptation to contrast. Perception 11:505-528. Rotshenker, S., Rahamimoff, R. (1970). Neuromsucular Synapse: stochastic propertes of spontaneous release of transmitter. Science 170:648-649. Ruseckas, J., Kaulakys, B. (2010). 1/f noise from nonlinear stochastic differential equations . arXiv:1002.4316v1 [nlin.AO] Sakai, Y., Funahashi, S., and Shinomoto, S. (1999). Temporally correlated inputs to leaky integrate -and-fire models can reproduce spiking statistics of cortical neurons. Neural Networks, 12:1181-1190. Salinas, E., and Sejnowski, T. J. (2002). Integrate-and-Fire Neurons Driven by Correlated Stochastic Input. Neural Computation, 14:2111-2155. Sales-Pardo, M.,Guimera, R., Moreira, A.A., Amaral, L.M.A (2007) Extracting the hierarchical organization of complex systems. Proc.Nat.Acad.Sci. USA 104(39): 15224-15229 Salvadori, G., Biella G. (1994). Discriminating properties of wide dynamic range Neurons my means of universal multifractals. In: Fractals in Biology and Medicine II, eds. G.A. Losa, D. Merlini, T.F. Nonnenmacher, E.R. Weibel,. pp. 314-325. Scafetta,N., Marchi,D., West, B.J. (2009). Understanding the complexity of human gait dynamics. Chaos 19:026108. Scheffer, M., Bascompte, J., Brock, W.A., Brovkin, V., Carpenter S.R., Dakos, V., Held, H., van Nes, E.H., Rietkerk., M.D.,Sugihara, G. Early-warning signals for critical conditions. Nature 461:53-59. Schierwagen, A. (2008). Neuronal Morphology: Shape Characteristics and Models. Neurophysiology, 40(4):310- 315. Segev, R., Benveniste, M., Hulata, E., Cohen, N., Palevski, A., Kapon, E., Shapira, Y., and Ben -Jacob, E. (2002). Long Term Behavior of Lithographically Prepared In Vitro Neuronal Networks. Physical Review Letters, 88(11), 118102. Segev, R., Baruchi, I., Hulata, E., and Ben-Jacob, E. (2004). Hidden Neuronal Correlations in Cultured Networks. Physical Review Letters, 92(11), 118102. Seuront, L. (2010). Fractals and Multifractals in Ecology and aquatic science. CRC Press, Boca Raton, Shadlen, M. N., and Newsome, W. T. (1998). The Variable Discharge of Cortical Neurons: Implications for Connectivity, Computation, and Information Coding. J Neurosci, 18(10):3870-3896. Shahverdian, A. Y., and Apkarian, A. V. (1999). On Irregular Behavior of Neuron Spike Trains. Fractals, 7(1):93- 103. Shew,W.L., Yang,H., Petermann T., Roy, R., Plenz, D. (2009). Neuronal avalanches imply maximum dynamic range in cortical networks at criticality. J. Neurosci. 29(49):15600-15595. Shimizu, Y., Barth, M., Windischberger, C., Moser, E., and Thurner, S. (2004). Wavelet -based multifractal analysis of fMRI time series. NeuroImage, 22:1195-1202. Shimono, M., Owaki, T., Aman, K., Kitajo, K., and Takeda, T. (2007). Functional modulation of power -law distribution in visual perception. Physical Review E, 75, 051902. Shinomoto, S., Shima, K., Tanji, J. (2003). Differences in Spiking Patterns Among Cortical Neurons. Neural Computation, 15:2823-2842. Shlesinger, M. F., West, B. J., and Klafter, J. (1987). Levy Dynamics of Enhanced Diffusion: Application to Turbulence. Physical Review Letters, 58(11):1100-1103. Simon, H. A. (1955). On a Class of Skew Distribution Functions. Biometrika, 42:425-440. Simon, H.A. (1962). The Architecture of Complexity. Proc. Amer. Philosoph. Soc. 106(6):467 -482. Simon, H.A. (1973).The organization of complex systems. In: Hierarchy Theory, edit. H.H. Pattee, Brazillier, NY, pp.3-27. Smith Jr., T. G., Marks, W. B., Lange, G. D., Sheriff Jr., W. H., and Neale, E. A. (1989). A fractal analysis of cell images. J Neurosci Methods, 27:173-180. Smolders, F. D. J., Folgering H. Th. M., (1977). Actions and interactions of CO2 and O2 on the controlling system of the lung ventilation. Thesis, Nijmegen University, Netherlands. Sokal, A., Bricmont, J. (2004) Defense of a modest Scientific Realism. In: Carroer, M., Roggenhofer, J., Kueppers, G., Banchard ,P. (eds). Knowledge and the World: beyond the Science wars. Springer New York Sokolov, I. M., Klafter, J., and Blumen, A. (2002). Fractional Kinetics. Physics Today, 48-54. Soltys, Z., Ziaja, M., Pawlinski, R., Setkowicz, Z., and Janeczko, K. (2001). Morphology of Reactive Microglia in the Injured Cerebral Cortex. Fractal Analysis and Complementary Quantitative Methods. J Neurosci. Res., 63:90-97. Song, C., Havlin, S., and Makse, H. A. (2005). Self-similarity of complex networks. Nature, 433:392-395. Song, C., Havlin, S., Makse, H.A. (2006). Orogins of gfractality in the growth of complex networks. Nature Physics 2:275-281. Sornette, D., (2000), Critical Phenomena in Natural Sciences. Springer, Berlin. Sporns, O., and Zwi, J. D. (2004). The Small World of the Cerebral Cortex. Neuroinformatics, 2:145-162. Sporns, O., Koetter, R. (2004) Motifs in Brain networks. PLoS Biology 2(11): e369 Sporns,O., Chialvo, D.R., Kaiser, M., Hilgetag, C.C. (2004). Organization, development and function of complex brain networks. Trends Cog. Sci. 8(9): 418-425. Sporns, O. (2006). Small-world connectivity, motif composition, and complexity of fractal neuronal composition. BioSystems, 85:55-64. Sporns, O., Honey, C.J., Koetter, R. (2007). Identification and classification of hubs in brain networks. PLoS ONE 10: e1049 Stam, C. J. (2004). Functional connectivity patterns of human magnetoencephalographic recordings: a ‘small -world’ network? Neurosci Letters, 355:25-28. Stam, C. J., and Bruin, E. A. d. (2004). Scale-Free Dynamics of Global Functional Connectivity in the Human Brain. Human Brain Mapping, 22:97-109. Stam, C. J. (2005). Nonlinear dynamical analysis of EE and MEG: Review of an emerging field. Clinical Neurophys., 116:2266-2301. Stam, C. J., and Reijneveld, J. C. (2007). Graph theoretical analysis of complex network in the brain. Nonlinear Biomedical Phys, 1:3. Stanley, H.E. (1987). Introduction to phase transitions and critical phenomena. Oxford University Press, Oxford, UK Stanley, H.E. (1999). Scaling, universality and renormalization: the three pillars of modern critical Phenomena. Rev. Mod. Physiscs 71:S358-S366. Stauffer,D., Aharony A. (1991/1994). Introduction to Percolation theory. CRC Press, Boca Raton. Stephen,D.G., Dixon, J.A. (2009) The self-organization of insight: Entropy and power laws in problem solving. J. of Problem Solving 2(1):72-101. Stevens, S. S. (1957). On the psychophysical Law. Psychol.Rev.64 (3):153-181. Stevens, C. F., and Zador, A. M. (1998). Input synchrony and the irregular firing of cortical neurons. Nature Neuroscience, 1(3):210-217. Stewart, C.V., Plenz, D. (2006) Inverted U profile of dopamine-NMDA mediated spontaneous avalanche recurrence in superficial layers o rat prefrontal cortex. J. Neurosci. 23:8148-8159. Stinchcombe, R. B. (1989). Fractals, phase transitions, and criticality. Proc. R. Soc. Lond. A, 423:17-33. Stoop, R., Wagner, C. (2007) Neocortex’s architecture optimizes computation, information transfer and synchronizability, at given total connection length. Int.J.Bifurc.Chaos 17(7): 2257-2279. Suckling,J., Wink A.M., Bernard, F.A., Barnes, A., Bullmore, E. (2008) Endogenous multifractal brain dynamics are modulated by age, cholinergic blockade and cognitive performance, J. Neuroscience Methods 174:292 -300. Tabor, W. (2000). Fractal encoding of context-free grammars in connectionist networks. Expert Systems 17:41-56. Takeda, T., Sakata, A., and Matsuoka, T. (1999). Fractal Dimensions in the Occurrence of Miniature End -Plate Potential In a Vertebrate Neuromuscular Junction. Prog. Neuro-Psychopharmacol & Biol. Psychiat., 23:1157-1169. Tateno, T., Kawana, A., Jimbo,Y. Analytical characterization of spontaneous firing in networks of developing rat cultured cortical neurons. Physical Review E 65:051924. Tebbens, S. F., and Burroughs, S. M. (2003). Self-Similar Criticality. Fractals, 11(3):221-231. Teich, M. C., and Saleh, B. E. A (1981). Interevent-time statistics for shot-noise-driven self-exciting point processes in photon detection. J. Opt. Soc. Am., 71(6):771-776. Teich, M. C., Johnson, D. H., Kumar, A. R., and Turcott, R. G. (1990). Ra te fluctuations and fractional power-law noise recorded from cells in the lower auditory pathway of the cat. Hearing Research, 46:41-52. Teich, M. C., Heneghan, C., Lowen, S. B., Ozaki, T., and Kaplan, E. (1997). Fractal character of the neural spike train in the visual system of the cat. . J. Opt. Soc. Am., 14(3):529-546. Teramae, J.-n., and Fukai, T. (2007). Local cortical circuit model inferred from power-law distributed neuronal avalanches. J Comput Neurosci, 22:301-312. Thatcher, R. W., North, D. M., and Biver, C. J. (2009). Self-Organized Criticality and the Development of EEG Phase Reset. Human Brain Mapping, 30:553-574. Thiagarajan T.C., Lebedev, M.A., Nicolelis M.A., Plenz, D.(2010), Coherence potentials: Loss -less, all-or-none network events in the cortex. PLoS Biology 8 (1):e1000278. Thorson, J., Biederman-Thorson, M. (1974). Distributed relaxation processes in sensory adaptation. Science 183:161-183. Thurner, S., Lowen, S. B., Feurstein, M. C., Heneghan, C., Feichtinger, H. G., and Teich, M. C. (1997). Analysis, synthesis, and estimation of fractal-rate stochastic point processes. Fractals, 5(4):565-595. Thurner, S., Windischberger, C., Moser, E., Walla, P., and Barth, M. (2003) Scaling laws and persistence in human activity. Physica A, 326:511-521. Tino, P. (1999). Spatial representation of symbolic sequences through iterative function systems. IEEE Trans. On Systems, Man and Cybernetics 29(4):386-393. Tognoli, E., Kelso, J.A.S. (2009). Brain Coordination Dynamics: True and false faces of phase synchrony and metastability. Progr.Neurobiol. 87:31-40. Toib, A., Lyakhov, V., and Marom, S. (1998). Interaction between Duration of Activity and Time Course of R ecovery from Slow Inactivation in Mammalian Brain Na+ Channels. J Neurosci, 18(5):1893-1903. Touboul, J., and Destexhe, A. (2009). Can power -law scaling and neuronal avalanches arise from stochastic dynamics? arXiv:0910.0805v1. Tsuda,I., Kuroda, S. (2004) A complex system approach to an interpretation of Dynamic Brain activity II. : does Cantor coding provide a dynamical model for the formation of episodic memory ? in: Cortical Dynamics, LNCS 3146, edit: P. Erdi, pp. 129-139. Springer, Berlin. Turcotte, d.L., (1999). Self-organized criticality. Rep.Prog.Phys. 62:1377-1429. Tsallis, C., Levy, S. V. F., Souza, A. M. C., and Maynard, R. (1995). Statistical -Mechanical Foundation of the Ubiquity of Levy Distributions in Nature. Physical Review Letters, 75(20):3589-3593. Tsallis, C. (2009). Introduction to non-extensive Statistical Mechanics: approaching a complex World . Springer, NY. Uhlhaas,P.J., Pipa, G., Lima, B., Melloni, L., Neuenschwander, S., Nikolic, D., Singer, W. (2009). Neural Synchrony in cortical networks : history, concept and current status. Frontiers in Integrative Neuroscience 3:1-19 Ulanovsky, N., Las, L., Farkas, D., Nelken, I. (2004). Multiple time scales of adaptation in auditory cortex neurons. J. Neurosci. 17:10440-10453 Usher,M., Stemmler, M., Koch,C., Olami,Z. (1994). Network amplification of local fluctuations causes high spike rate variability, fractal firing patterns and oscillatory local field potentials. Neural Computation 6:795-836. Usher, M., and Stemmler, M. (1995). Dynamic Pattern Formation Leads to 1/f Noise in Neural Populations. Physical Review Letters, 74(2):326-329. Van den Heuvel, M. P., Stam, C. J., Boersma, M., and Pol, H. E. H (2008). Small -world and scale-free organization of voxel-based resting-state functional connectivity in the human brain. NeuroImage, 43:528-539. Van Order, G. C., Holden, J. G., and Turvey, M. T. (2003). Self -Organization of Cognitive Performance. J Experimental Psych: General, 132(3):331-350. Van Pelt, J., Coner, M.A., Wolters, P.S., Rutten, W.L.C., Ramakers, G.J.A. (2004). Longterm stability and developmental changes in spontaneous network burst firing patterns in dissociated rat cerebral cortex cell cultures on microelectrode arrays. Neuroscience Letters 361:86-89. Van Vreeswijk, C. (2001). Informat ion transmission with renewal neurons. Neurocomputing, 38-40:417-422. Varanda, W. A., Liebovitch, L. S., Figueiroa, J. N., and Nogueira, R. A. (2000). Hurst Analysis Applies to the Study of Single Calcium-activated Potassium Channel Kinetics. J. theor. Biol., 206:343-353. Varela F., Lachaux, J-P., Rodriguez,E., Martinerie,J. (2001), The Brainweb: phase synchronization and large scale integration. Nature Reviews Neuroscience 2:229-239. Voss, R. F., and Clarke, J. (1975). „1/f noise‟ in music and speech. Nature, 258:317-318. Vrobel, S. (2007) Fractal time, observer perspective and levels of description in Nature.. Electronic J. of Theoretical Physics 16(II):275-302. Wagenaar, D.A., Pine, J., Potter, S.M. (2006 a) An extremely rich repertoire of bursting pa tterns of bursting patterns during the development of cortical cultures. BMC Neuroscience 7: 7-11. Wagenaar, D.A.,Nadasky, Z., Potter, S.M. (2006 b). Persistent dynamic attractors in activity patterns of cultured neuronal networks. Physical Review E 73: 051907. Wagenmakers, E.-J., Farrell, S., and Ratcliff, R. (2004). Estimation and interpretation of 1/f α noise in human cognition. Psychonomic Bulletin & Review, 11(4):579-615. Wagenmakers, E-J., Farrell, S. and Ratcliff, R. (2005). Human Cognition and a pile of sand: a discussion on serial correlations and self-organized criticality. J.Exp.Psychol.Gen. 134(1): 108-116. Wark, B., Lundstrom, B.N., Fairhall, A. (2008). Sensory Adaptation. Curr. Opin. Neurobiol. 17:423-429. Watts, D., Strogatz S. (1998). Collective dynamics of Small World Networks. Nature 393:440-442. Wen, Q., Stepanyants, A., Elston, G. N., Grosberg, A. Y., and Chklovskii, D. B. (2009). Maximization of the connectivity repertoire as a statistical principle governing the shapes of dendritic arbors. Proc.Natl.Acad.Sci. USA, 106:12536-12541. West, B.J., Griffin, L. (1999).Allometric control, Inverse power laws and human gait. Chaos, Solitons & Fractals 10(9):1519-1527. Werner, G., and Mountcastle, V. B. (1963). The Variability of Central Neural Activity in a Sensory System, and its Implications for the Central Reflection of Sensory Events. J Neurophysiol., 26:958-977. Werner, G., and Mountcastle, V. B. (1964). Neural Activity in Mechanoreceptive Cutaneous Afferen ts; Stimulus- Response Relations, Weber Functions, and Information Transmission. J. Neurophysiol., 28:359-394. Werner, G. (2007a). Perspectives on the Neuroscience of Cognitions and Consciousness. BioSystems, 87:82-95. Werner, G. (2007b). Metastability, criticality, and phase transitions in brain and its models. BioSystems, 90:496-508. Werner, G. (2009a). Viewing brain processes and Critical State Transitions across levels of organization: Neural events in Cognition and Consciousness, and general principles. BioSystems, 96:114-119. Werner, G. (2009b). Consciousness related neural events viewed as brain state space transitions. Cogn. Neurodyn., 3:83-95. Werner, G. (2009c). On Critical State Transitions Between Different Levels in Neural Systems. New Math. And Natural Computation, 5(1):185-196. West, B.J. (1999 a). Physiology, Promiscuity and Prophecy at the Millenium: a tale of tails . Studies of Nonlinear Phenomena in the Life Sciences, Vol. 7. World Scientific, Singapore. West, B.J. (2006). Where Medicine went wrong: rediscovering the path to complexity. World Scientific,Singapore. West, B.J. (2009 b). Control from an allometric perspective . Advances in Experimental Medicine and Biology 629:57-82. West, B. J., Allegrini, P., Grigolini, P. (1994). Dynamical Generators of Levy Statistics in Biology. In: Fractals in Biology and Medicine, Vol. II, eds. G.A. Losa, D. Merlini, T.F. Nonnenmacher, E.R. Weibel. Birkhauser, Basel. West, B.J., Deering, B. (1995). The Lure of Modern Science. Studies in Nonlinear Phenomena in Life Sciences, Vol. 3. World Scientific. West, B. J., Grigolini, P., Metzler, R., and Nonnenmacher, T. F. (1997). Fractional diffusion and Levy stable processes. Physical Review E, 55(1):99-106. West, B. J., and Griffin, L. (1999 a). Allometric Control, Inverse Power Laws and Human Gait. Chaos, Solitons, and Fractals, 10(9):1519-1527. .West, B. J., and Nonnenmacher, T. (2001). An ant in a gurge. Physics Letters A, 278:255-259. West, B.J., Bologna, M., Grigolini, P. (2003). Physics of Fractal Operators. Springer. West, B.J., Scafetta N. (2003) Nonlinear dynamical model of human gait. Physical Review E 67:051917. West, B. J., Geneston, E. L., Grigolini, P.(2008). Maximizing Information Exchange between complex Networks. Physics Reports, 468(1-3), 1-99. Willinger,W., Taquu, M.S., Sherman, R., Wilson, D.V., (1995). Self-similarity through high variability: statistical analysis of Ethernet LAN traffic at the source level. ACM SIGCOMM Computer Communication Review 25 (4):100- 113. Willinger, W. (2000). The Discovery of Self-Similar Traffic. Lecture Notes in Computer Science 1769. Springer. pp 513-527. Wilson, K. G. (1979). Problems in Physics with Many Scales of Length. Sci. Amer., 241:158-179. Wise, M.E. (1981). Spike Interval Distributions for Neurons and Random Walks with Drift to a Fluctuating Threshold. Statistical Distributions in Scientific Work, Taillie et al, eds., 6:211-231, Reidel Publ. Comp. Witten Jr., T. A., and Sander, L. M. (1981). Diffusion-Limited Aggregation, a Kinetic Critical Phenomenon. Physical Review Letters, 47(19):1400-1403. Womelsdorf, T., Schoffelen,J.M., Oostenveld, R., Singer W., Desimone,R., Engel,A,K, Fries,P. (2007). Modulation of neuronal interactions through neuronal synchronization. Science 316:1609-1612. Wornell, G. W. (1993). Wavelet-Based Representations for the 1/f Family of Fractal Processes. Proceedings of the IEEE, 81(10):1428-1450. Xu, Z., Payne, J. R., and Nelson, M. E. (1996). Logarithmic Time Course of Sensory Adaptation in Electrosensory Afferent Nerve Fibers in a Weakly Electric Fish. J Neurophysiol., 76(3):2020-2032. Yamamoto, M., Nakahama, H., Shima, K., Kodama, T., Mushiake, H. (1986). Markov dependency and spectral analyses on spike counts in mesencephalic reticular neurons during sleep and attentive states. Brain Research 366:279-289. Yu, Y., Romero, R., Lee,T.S. (2005). Preference of sensory neural coding for 1/f signals. Phys. Rev. Lett. 94:108103. Yule, G. U. (1925). A Mathematical Theory of Evolution, based on the Conclusions of Dr. J. C. Willis, F. R. S. Phil. Trans. R. Soc. Lond. B, 213:21-87. Zanette, D. H. (2008). Zipf‟s law and the creation of musical context. arXiv:cs/0406015v1. Zapperi, S., Lauritsen, K.B., Stanley, H.E. (1995). Self-organized branching process: mean-field theory for avalanches. Physical Review Letters 75(22): 4071-4074. Zarahn, E., Aguirre, G. K., and D‟Esposito, M. (1997). Empirical Analyses of BOLD fMRI Statistics. NeuroImage, 5:179-197. Zhang, L. (2006).Quantifying brain white matter structural changes in normal aging using fractal Dimension. Doctoral Thesis, Case Western Reserve University. http://etd.ohiolink.edu/send- pdf.cgi/Zhang%20Luduan.pdf?acc_num=case1126213038 Zietsch, B. and Elston, G. N. (2005). Fractal Analysis of Pyramidal Cells in the Visual Cortex of the Galago (Otolemur Garnetti): Regional Variation in Dendritic branching Patterns Between Visual Areas. Fractals, 13(2):83- 90. Zhou, C., Zemanova, L., Zamora-Lopez, G., C. C., and Kurths, J. (2007). Structure-function relationship in complex brain networks expressed by hierarchical synchronization. New Journal of Physics, 9, 178. Zumofen, G., and Klafter, J. (1993). Scale-invariant motion in intermittent chaotic systems. Physical Review E, 47(2):851-863.
1802.08279
4
1802
2019-03-20T16:04:03
Implementing a Concept Network Model
[ "q-bio.NC" ]
The same concept can mean different things or be instantiated in different forms depending on context, suggesting a degree of flexibility within the conceptual system. We propose that a compositional network model can be used to capture and predict this flexibility. We modeled individual concepts (e.g., BANANA, BOTTLE) as graph-theoretical networks, in which properties (e.g., YELLOW, SWEET) were represented as nodes and their associations as edges. In this framework, networks capture the within-concept statistics that reflect how properties correlate with each other across instances of a concept. We ran a classification analysis using graph eigendecomposition to validate these models, and find that these models can successfully discriminate between object concepts. We then computed formal measures from these concept networks and explored their relationship to conceptual structure. We find that diversity coefficients and core-periphery structure can be interpreted as network-based measures of conceptual flexibility and stability, respectively. These results support the feasibility of a concept network framework and highlight its ability to formally capture important characteristics of the conceptual system.
q-bio.NC
q-bio
1 The updated and peer-reviewed version of this paper is now published. Please read and refer to this final version: Solomon, S.H., Medaglia, J.D., & Thompson-Schill, S.L. (2019). Implementing a concept network model. Behavior Research Methods. https://doi.org/10.3758/s13428-019- 01217-1 2 Implementing a Concept Network Model Sarah H. Solomon1, John D. Medaglia2,3, and Sharon L. Thompson-Schill1 1 Department of Psychology, University of Pennsylvania 2 Department of Psychology, Drexel University 3 Department of Neurology, Perelman School of Medicine, University of Pennsylvania May 22, 2018 Abstract The same concept can mean different things or be instantiated in different forms depending on context, suggesting a degree of flexibility within the conceptual system. We propose that a compositional network model can be used to capture and predict this flexibility. We modeled individual concepts (e.g., BANANA, BOTTLE) as graph-theoretical networks, in which properties (e.g., YELLOW, SWEET) were represented as nodes and their associations as edges. In this framework, networks capture the within-concept statistics that reflect how properties correlate with each other across instances of a concept. We ran a classification analysis using graph eigendecomposition to validate these models, and find that these models can successfully discriminate between object concepts. We then computed formal measures from these concept networks and explored their relationship to conceptual structure. We find that diversity coefficients and core-periphery structure can be interpreted as network-based measures of conceptual flexibility and stability, respectively. These results support the feasibility of a concept network framework and highlight important its ability characteristics of the conceptual system. Keywords: conceptual knowledge, conceptual flexibility, network science to formally capture 3 Introduction The APPLE information evoked by "apple pie" is considerably different from that evoked by "apple picking": the former is soft, warm, and wedge-shaped, whereas the latter is firm, cool, and spherical. If you scour your conceptual space for APPLE information, you will uncover the knowledge that apples can be red, green, yellow, or brown when old; that they can be sweet or tart; that they are crunchy when fresh and soft when baked; that they are naturally round but can be cut into slices; that they are firm, but mushy if blended; that they can be found in bowls, in jars, and on trees. Despite the complexity of this conceptual knowledge, we can generate an appropriate APPLE instance, with the appropriate features, based on the context we are in at the time. In other words, the multi-faceted APPLE concept can be flexibly adjusted in order to enable a near-infinite number of specific and appropriate APPLE exemplars. How is conceptual knowledge structured such that this generative and essential flexibility is possible? In particular, we are interested in the structure of individual concepts (e.g., APPLE, SNOW), rather than the structure of super-ordinate categories (e.g., FRUIT, TOOLS) or the structure of semantic space more broadly. This latter pursuit -- the modeling of semantic space -- has already been approached from various theoretical orientations and methodologies. In "compositional" cognitive theories, the meaning of a concept can be decomposed into features and their relationships with each other (e.g., Smith et al, 1974, McRae et al., 1997; Tyler & Moss, 2001). This compositionality can be incorporated into the associated computational models: researchers primarily use various forms of feature-based connectionist models, in which concepts are represented as patterns of activation over features, to simulate semantic behavior (e.g., Cree et al., 1999; 2006; Randall et al., 2004). In "relational" frameworks, concepts are defined in terms of how they relate to other concepts in semantic or lexical space (e.g., Landauer & Dumais, 1997). Researchers within this framework can use word co-occurrence or association statistics to create large semantic networks, which can be analyzed using a rich set of network science tools (e.g., Steyvers & Tenenbaum, 2005; Van Rensbergen et al., 2015; De Deyne et al, 2016). The compositional approach to conceptual knowledge generally represents individual concepts as vectors of features. These features can span a range of 4 information-types (e.g., visual, functional, encyclopedic), consistent with a distributed account of conceptual knowledge. The "conceptual structure" account (Tyler & Moss, 2001) represents concepts as binary vectors indicating the presence of absence of features, and argues that broad semantic domains (e.g., ANIMALS, TOOLS) differ in their characteristic properties and in their patterns of property-correlations (e.g., HAS-WINGS and FLIES tend to co-occur within the ANIMAL domain). The "feature-correlation" account (e.g., McRae et al., 1997; 1999; McRae, 2004) tweaks this model by empirically deriving conceptual property statistics and by implementing this framework in a type of connectionist model called an attractor network: property statistics characterize the structure of conceptual space, and the model can leverage these statistics to settle on an appropriate conceptual representation given the current inputs (Cree et al., 1999; Cree et al., 2006). These models contain a dynamic component in that the attractor networks reveal how word comprehension may unfold over time, and is "flexible" in the sense that the model may follow varied trajectories through conceptual space in order to settle on a specific concept's representation. However, the ability of this approach to capture conceptual flexibility -- in the sense described above -- is not fully fleshed out. A feature-based framework is valuable because the elements that can be adjusted during conceptual processing are explicitly modeled, but a different set of tools may be helpful in the pursuit of capturing how flexibility might emerge out of this conceptual structure. Our initial steps to develop such tools are reported here. Another way to model conceptual knowledge is to capture statistical relations between words or phrases in language. This approach is "relational", rather than compositional, because a concept's meaning is represented in terms of its relations to other concepts, rather than assuming any kind of internal conceptual structure. Word co-occurrence statistics can be extracted from text corpora and have been used to create probabilistic models of word meanings (Griffiths et al., 2007), to represent semantic similarity (Landauer & Dumais, 1997), and to characterize the structure of the entire lexicon (e.g., WordNet; Miller & Fellbaum, 2007). In a similar approach, word association data can be used to capture and analyze the structure of semantic space (Steyvers & Tenenbaum, 2005; Van Rensbergen et al., 2015; De Deyne et al., 2016). These data are generally modeled as networks, which can be analyzed in formal ways. 5 The use of networks to model semantic knowledge has a well-established history. The early "semantic network" models (Collins & Quillian, 1969; Collins & Loftus, 1975) represent concepts as nodes in a network; links between these nodes signify associations in semantic memory. These networks capture the extent to which concepts are related to other concepts and features, and can model the putatively hierarchical nature of conceptual knowledge. Though these models are "network-based", they are so in a rather informal way. On the other hand, network science, a mathematical descendent of graph theory, has developed a rich set of tools to study networks in a formal, quantitative framework (Barabási, 2016). There has been a recent surge of research applying network science to neural and linguistic data, providing new insights into these complex systems along with new tools and measures we can use to study them. Formal networks are composed of units (i.e., "nodes") and the links between them (i.e., "edges"). Current network science approaches to semantic and lexical knowledge use nodes to represent individual words, and edges to represent their co-occurrence or association statistics. Once modeled in this way, aspects of network structure can be quantitatively analyzed and relationships between network structure and other phenomena can be explored. For example, it has been suggested that human language exhibits small-world properties (Steyvers & Tenenbaum, 2005; i Cancho & Sole, 2001), and that semantic networks exhibit an "assortative" structure, meaning that semantic nodes tend to have connections to other semantic nodes with similar characteristics (e.g., valence, arousal, concreteness; Van Rensbergen et al., 2015). A spreading activation model applied to these word-association networks makes accurate predictions of weak similarity judgments (De Deyne et al., 2016). Further, Steyvers & Tenenbaum (2005) report that a word's degree (i.e., how many links it has to other nodes) predicts both age of acquisition and reaction times on lexical decision tasks. The application of network science tools to semantic data enables researchers to explore higher-level structure in semantic space, and to use these structural characteristics to predict aspects of semantic and lexical processing. However, this approach does not examine the internal structure of concepts: individual concepts are characterized in terms of their statistical co-occurrences with other concepts in language, and not in terms of their 6 Figure 1: Example images used to generate test data in classification analysis. Test data used in the classification analysis were generated from participants who made property judgments on images of conceptual exemplars. Yellow cross indicates object to be considered. Example images for grass (top) and cookie (bottom). unique, internal content. That is, these language-based approaches are non- compositional, and we argue that compositionality is a key aspect of flexible conceptual models. Modeling a concept's internal structure -- along with its features and the ways those features interact -- enables us to flexibly adjust which features are included, and to what degree, in a given conceptual instance. We believe that a compositional conceptual framework paired with network science techniques provides a platform on which to model conceptual flexibility. Unlike previous network approaches, we use networks to model individual concepts, rather than the semantic system as a whole. In this case, the nodes in each concept's network represent individual features, and the edges of the network (i.e., the links between the nodes) represent the statistical relationship between features within that concept. That is, edges capture the extent to which certain properties tend to covary with each other within a concept. The creation of such networks thus depends on our ability to calculate within-concept statistics. These statistics provide the scaffolding to build our networks, and also reveal how a concept's information may be appropriately adjusted to form valid, yet varied, instances of that concept. Our goal is to show that creation of such networks is possible, and that they can be used to capture conceptual flexibility, and other phenomena, in a formal way. We chose 15 basic-level concepts (e.g., CHOCOLATE, TABLE, GRASS, KNIFE) and defined a list of within-concept states for each one (e.g., DARK 7 CHOCOLATE, WHITE CHOCOLATE, CHOCOLATE SYRUP, CHOCOLATE CHIPS) using a large sample of participants on Amazon Mechanical Turk (AMT). We then compiled a large set of conceptual properties that could apply to any of the 15 concepts (e.g., BROWN, GREEN, WOODEN, METAL, SHARP, SWEET). A final sample of AMT participants then reported which of the properties corresponded to each of the specific concept-states. This let us know which properties applied to dark chocolate (e.g., BLACK, BITTER) and which applied to white chocolate (e.g., WHITE, SWEET). These data enable us to calculate the within-concept statistics necessary to construct our networks. Each of the properties corresponds to a vector that denotes whether that property is present or absent in each of the concept-states, and we can correlate these property vectors with each other to determine the extent to which each property covaries with every other property within that specific concept. These correlation values are encoded as the edges in the concept's network. Once these networks were constructed, we had two objectives: First, we intended to determine whether or not these networks contain concept- specific information. We thus ran a classification analysis over these networks to confirm that within-concept statistics can be used to discriminate between concepts. We performed eigendecomposition on our concept networks in a classification analysis, which provided a measure of the extent to which a vector is consistent with an underlying network structure (e.g., Medaglia et al., 2017, Huang et al., in press). In our case, our test data were vectors of properties generated from photographs of individual concept exemplars (Fig. 1). Second -- and most importantly -- we intended to extract useful and interpretable measures from our concept networks. In particular, we aimed to quantify conceptual flexibility. Many networks by their very nature permit flexibility, because a single network can support different states, each characterized by different patterns of activation across nodes. A node's contribution to network flexibility can be understood in terms of its position in the context of the larger network. Most natural systems exhibit "small-world" network structure (Bassett & Bullmore, 2017), which means that there are clusters of nodes in a network with strong connections between them (Amaral et al., 2000). These are called "modules", and nodes can interact with these modules in different ways. Some nodes may have links that are highly distributed across the modules in a network, whereas other nodes may have links only in one module. Each node in a network can be assigned a diversity coefficient, a 8 version of the participation coefficient calculated using normalized Shannon entropy, which reflects this tendency. We interpreted network diversity as a likely candidate for a formal flexibility measure, and pooled the diversity coefficients across nodes in a concept's network to quantify that concept's flexibility. We calculated this version of flexibility for our 15 basic-level concepts, and predicted that it would correlate with a measure of "semantic diversity" calculated separately using word co-occurrence statistics (SemD; Hoffman et al., 2013). This would suggest that network-based measures can successfully be used to quantify flexibility in a compositional conceptual framework. Another phenomenon of interest to cognitive scientists is the distinction between context-independent and context-dependent conceptual properties (Barsalou, 1982). Context-independent properties are those that are automatically activated for a concept in all contexts, and are sometimes referred to as "core" properties. On the other hand, context-dependent properties are those that are only activated when the context renders them relevant. Concepts are composed of both kinds of properties, such that some properties are stable and are activated across all instances, and some are more variable and are only activated some of the time. The distinction between context-independent and -- dependent properties has been suggested in reaction time differences in property-verification tasks (Barsalou, 1982), but the classification of a property as one type or the other has been decided upon by the experimenter rather than being calculated in a quantitative way. One of our goals was to use our concept networks to extract this information; that is, whether the structure of each concept does in fact include such a core. Network science provides techniques for assessing this core-periphery structure (Borgatti & Everett, 2000; Bassett et al., 2013). In network terms, a core is a set of nodes that are densely interconnected and therefore often co- activated, whereas the periphery consists of nodes with sparser connections. A measure can be extracted that represents the extent to which a given network has a core-periphery structure; some networks might have more prominent cores than others. We hypothesize that this construct of core- periphery structure can provide a way to formally capture the notion of context-dependent and context-independent conceptual properties. In other words, perhaps a certain concept has a large set of context-independent properties that are consistently activated across a large range of contexts: this concept's network might have a strong core-periphery structure. It also 9 seems reasonable to suggest that those concepts with a stronger core might be less flexible in the ways described above. If we interpret a core as a set of properties whose activation patterns are stable across contexts, then there is less room for variability in the expression of those properties, and therefore less flexibility overall. On the other hand, more flexible concepts might have a weaker core-periphery structure, reflecting the more variable patterns of property activations. We therefore predict a negative relationship between our network measures of flexibility and core-periphery structure. Methods General Methods Network Construction In order to create our networks we first had to define our nodes. Since our nodes represent individual conceptual properties, we compiled a list of properties that could be applied to all of our target concepts. Participants were recruited from Amazon Mechanical Turk (AMT) and were asked to list all of the properties that must be true or can be true for each concept. It was emphasized that the properties do not have to be true of all types of the concept. Participants were required to report at least 10 properties per concept, but there was no limit on the number of responses they could provide. Once these data were collected, we organized the data as follows. For each concept, we collapsed across different forms of the same property (e.g., "sugar", "sugary", "tastes sugary"), and removed responses that were too general (e.g., "taste", "color"). For each concept, we only included properties that were given by more than one participant. We then combined properties across all concepts to create our final list of N properties that will be represented as nodes in our concept networks. The same AMT participants that provided conceptual properties also provided concept-states for each of the target concepts. For each concept, participants were asked to think about that object and all the different kinds, forms, types, or states in which that object can be found. Participants were required to make at least five responses, and could make up to 15 responses. For each concept, we removed responses that we considered properties rather than types (e.g., "sweet chocolate"), and responses that were too specific (e.g., "Chiquita banana"). We only included responses that were 10 given by more than one participant, resulting in a set of K concept-states for each concept. A separate set of AMT participants was presented with one concept-state of each of the target concepts in random order (e.g., "dark chocolate", "frozen banana") and was asked to select the properties that are true of that specific concept-state. The full list of N properties was displayed in a multiple-choice format. For each concept-state, responses were combined across participants and represented in a binary fashion. A property was only considered "true" for a concept-state if more than one participant made that response. At this point, each concept's data included a set of K concept-states, each of which corresponds to a N-length vector that indicates the presence or absence of each property. We could also view these data as a set of N conceptual properties, each of which corresponded to a K-length vector that indicates its presence or absence in each of the concept-states. For each concept, we excluded properties that were not present in any of the concept-states, resulting in a smaller set of NC properties. We created a network by correlating the NC binary property-vectors with each other to create a NC x NC symmetrical, weighted correlation matrix. The diagonal was set to 0, and these networks were filtered using the triangulation filtering method (Massara et al., 2016; Tumminello et al., 2005; Kenett et al., 2014). This filtering approach generates a simpler subgraph that maximizes information content while reducing the influence of noise. This method is appropriate for graphs where edges are defined as correlations between nodes, as is the case here. No parameter fitting is required. These final, filtered concept networks were then analyzed using standard network science methods. Classification Analysis If our concept network models capture concept- specific information, the networks should be able to successfully discriminate between new concept exemplars. Exemplar data were generated from sets of photographs for each concept; all concept-states were represented. AMT participants were shown one image per concept, were asked to imagine interacting with this object in the real world, and to consider what properties it has. The full list of N properties was displayed in multiple-choice format, and participants were asked to select the properties that they believed applied to the object in the image. Individual participants' responses to each concept-state were represented as N-length property vectors and were used as test data in the classification analysis. 11 𝑥" = ∑ (𝑣(∙𝑥)+ ,(-. (1) By performing eigendecomposition on each adjacency matrix (i.e., concept network) we can assess the extent to which a vector is expected given an underlying network structure (e.g., Medaglia et al., 2017; Huang et al., in press). For each adjacency matrix A, V is the set of NC eigenvectors, ordered by eigenvalue. M is the number of ordered eigenvectors to include in analysis, and designates a subset of V. For each eigenvector v, we find the dot product with signal vector x, which gives us the projection of x on that dimension in the eigenspace of A. That is, it gives us an "alignment" value for that particular signal and that particular eigenvector. We can include all eigenvectors in M by taking the sum of squares of the dot products for each eigenvector. The alignment value for each signal is defined as: where 𝑥 is a property vector, M is the number of eigenvectors to include in alignment (sorted by eigenvalue), 𝑣( is one of M eigenvectors of the adjacency matrix, and 𝑥" is the scalar alignment value for signal x with The concept network that resulted in the highest alignment value (𝑥") was adjacency matrix A, given the eigenvectors 1-M. In our case, signal x is a property vector corresponding to a particular exemplar image (e.g., Fig. 1), which we align with each of the concept networks. Each exemplar was restricted to the properties included in each concept model before transformation; that is, exemplar data (x) were reduced to NC -- length vectors. taken as the "guess" of the classifier; each exemplar was either classified correctly (1), or incorrectly (0). We averaged these data across all exemplars to calculate the average classifier accuracy. To calculate a baseline measure of classification accuracy, we created traditional vector models for each concept. These models were similar to those used elsewhere in the literature (Tyler & Moss, 2001; McRae et al. 1997; 1999; 2004). For each concept, we averaged the K concept-state vectors resulting in an NC -length vector containing mean property strength values. Each concept's traditional vector model and network model contained the same conceptual properties. We ran a separate classification analysis using these traditional models and a correlational classifier. Each exemplar property-vector was correlated with each of the traditional concept vector models; the concept model that resulted in the highest correlation value was taken as the guess of the classifier. We calculated average 12 measures of classifier performance using the same methods described above, and also calculated classification accuracy within each concept. Network Analysis We extracted network metrics from our concept networks using the Brain Connectivity Toolbox (Rubinov & Sporns, 2010). The set of nodes in each network is designated as N, and n is the number of nodes. The set of links is L, and l is the number of links. The existence of a link between nodes (i,j) is captured in 𝑎(0: 𝑎(0=1 if a link is present and 𝑎(0=0 if a link is absent. The weight of a link is represented as 𝑤(0, and is normalized such that 0≤𝑤(0≤1. 𝑙6 is the sum of all weights in the degree, modularity (𝑄), core-periphery structure, and diversity coefficients connections to other nodes. Node degree (𝑘) is the number of connections Sporns, 2010). In weighted (i.e., non-binary) networks, node strength (𝑘6) is Nodes within a network differ in the number and strength of their that each node has with other nodes in the network (Eq. 2; Rubinov & calculated by summing the weights of the connections with other nodes (Eq. 3; Rubinov & Sporns, 2010). We separately averaged node strength and node degree within each network to obtain mean strength and degree measures for each concept network. We also counted the number of edges in each network overall. network. The network metrics we extracted included node strength, node (Fig. 2). 𝑘( = ∑ 𝑎(0 0∈: 𝑘(6 = ∑ 𝑤(0 0∈: We can attempt to partition a weighted network into sets of non-overlapping nodes (i.e., modules) such that within-module connections are maximized and between-module connections are minimized. Some networks exhibit modularity for each weighted network (Eq. 4; Rubinov & Sporns, 2010), which is defined as Modularity (𝑄) is a metric that describes a network's community structure. more of a modular structure than others; 𝑄6 is a quantitative measure of 𝑄6 = .;< ∑ (,0∈: =𝑤(0− ?@<?A<;< B 𝛿D@,DA (4) (2) (3) 13 Figure 2: Schematics of network structure. (A) Low-modularity network that contains nodes with equal degree. (B) High-modularity network with nodes in either module 1 (red) or module 2 (blue). One node (purple) participates in both modules; this is a high-diversity node. (C) Network with a strong core- periphery structure; some nodes comprise a densely connected core (purple) and others a weakly connected periphery (grey). where 𝛿D@,DA=1 if nodes i,j are in the same module (m), 𝑤(0 is the specific strength between nodes i,j, and ?@<?A<;< scales 𝑤(0 by the total strengths of the set of modules (M): Nodes may have connections to many different (ℎ(±) is a measure ascribed to individual nodes that reflects the diversity of nodes i,j across the network. Given a network's community structure, we can observe how individual nodes participate with each of the modules in modules, or have very few such connections. The diversity coefficient connections that each node has to modules in the network. This is a version of the participation coefficient, and is calculated using normalized Shannon entropy; we have previously used entropy to model property flexibility, and so predicted that diversity would be a good candidate for a network-based measure of conceptual flexibility. The diversity coefficient (Eq. 5; Rubinov & Sporns, 2011) for each node is defined as ℎ(± = − .HIJD∑ LÎ, Q@± , 𝑠(±(𝑢) is the strength of node 𝑖 within module 𝑢, and where 𝑝(±(𝑢) = Q@±(L) 𝑚 is the number of modules in modularity partition 𝑀. We averaged (𝑢)log𝑝(±(𝑢), diversity coefficients across nodes in a network to obtain a mean measure of diversity for each concept network. Core-periphery structure is another way to describe the structure of a network. Here, we attempt to partition a network into two non-overlapping 𝑝(± (5) (6) ∑ (,0ÎV_ Z𝑤(0−𝛾V𝑤\] ' sets of nodes such that connections within one set are maximized (i.e., the "core") and connections in the other are minimized (i.e., the "periphery"). can be partitioned in this way (Eq. 6), and can be defined as 14 Core-periphery fit (𝑄V) is a quantitative measure of how well each network 𝑄V= .WX Y∑ (,0ÎV^ where 𝐶b is the set of all nodes in the core, 𝐶cis the set of nodes in the periphery, 𝑤\ is the average edge weight, 𝛾Vis a parameter controlling the size of the core, and 𝑣Vis a normalization constant (Rubinov et al., 2015). Z𝑤(0−𝛾V𝑤\] − Methods: Set 1 The 5 concepts used in Set 1 were CHOCOLATE, BANANA, BOTTLE, TABLE, and PAPER. AMT participants (N=66) provided general properties for each concept along with concept-states. Another group of AMT participants (N=198) made property judgments on specific concept-states, and another group of AMT participants (N=60) generated test data for the classification analysis by making property judgments on individual images. The final property list included 129 properties. The number of states for each concept were as follows: chocolate=14, banana=15, bottle=11, table=14, paper=20. The full list of states can be seen in the Appendix. In the classification analysis, test data comprised a total of 300 property-vectors, with 60 exemplars/concept. Methods: Set 2 The 10 concepts used in Set 2 were KEY, PUMPKIN, GRASS, COOKIE, PICKLE, KNIFE, PILLOW, WOOD, PHONE, and CAR. AMT participants (N=60) provided general properties for each concept along with concept-states. Another group of AMT participants (N=108) made property judgments on specific concept- states, and another group of AMT participants (N=30) generated test data for the classification analysis by making property judgments on individual images. The final property list included 276 properties. The number of states for each concept were as follows: key=19, pumpkin=18, grass=16, cookie=22, pickle=17, knife=15, pillow=16, wood=22, phone=16, car=20. The full list of states can be seen in the Appendix. In the classification analysis, test data comprised 300 property-vectors, with 30 exemplars/concept. 15 eigen-dimensions from our concept networks. Classification was successful using ≥ 7 dimensions in Set 1, and ≥1 dimension in Set 2. Classification performance increased as more dimensions were added, such Figure 3: Classification results. We ran a range of classification analyses using different numbers of that performance of the network-models approached performance of the vector-based models (single data points). The sharp increase in performance in both sets is driven by eigenvectors with eigenvalues of 0, suggesting that contribution of individual features, suggesting that the presence Results Classification Results In order to determine whether our concept networks contained concept- specific information, we ran a classification analysis using eigendecomposition for both Set 1 and Set 2. We ran multiple analyses using different ranges of eigenvectors, which were sorted by eigenvalue (positive to negative). We started by only using the first eigenvector in each of the concept networks and determined whether this dimension alone could be used to classify the property vector. One dimension was enough to classify exemplars in Set 2 (Mean Accuracy=0.27; SE=0.03; Chance=0.10) but not Set 1 (Mean Accuracy=0.11; SE=0.02; Chance=0.20). However, increasing the number of dimensions improved classification performance for both sets (Fig. 3): for example, classification performance is significantly above chance when only 10 dimensions are used in Set 1 (Mean Accuracy=0.38; SE=0.03; Chance=0.10) and Set 2 (Mean Accuracy=0.38; SE=0.03; Chance=0.20). As more dimensions were 16 Table 2: Correlation results. We analyzed relationships between cognitive variables (SemS, SemD), network variables (modularity, core fit, mean diversity, mean strength, mean degree, number of edges) and vector-based classification results. *: p<0.05, **: p<0.01. included in the analysis, classification performance significantly increases and approaches the performance of the vector-based classifier, which was successful at classifying exemplars in Set 1 (Mean Accuracy=0.85; SD=0.06; Chance=.20) and Set 2 (Mean Accuracy=0.84; SD=0.10; Chance=0.10). The middle range on the x-axis of Fig. 3 contains eigenvectors with eigenvalues of 0: this is a special case in which a single property is weighted as 1. When one of these eigenvectors is multiplied by a signal (i.e., property vector) that includes that particular property, the alignment value is driven by the presence of that one property. In other words, the dramatic increase in classification performance in both Set 1 and Set 2 is driven by the successive contributions of individual features moving from left to right, suggesting that the presence or absence of individual features is highly informative for discriminating between concepts. Nevertheless, the significant classification performance using eigenvectors representing multiple features does suggest that our concept networks contain concept-specific information, motivating us to look within a concept for structural elements that relate to conceptual flexibility. It is this main goal that we pursue in the subsequent analyses. Network Measures of Conceptual Structure Networks across the two sets differed in node assignments, since they were constructed using different properties. However, once classification and network measures were extracted, we could pool the concepts together (N=15) and examine relationships between these network-related measures and other variables of interest. 17 Figure 4: A network-based measure of conceptual flexibility. Semantic diversity measures calculated using word co-occurrence statistics (Hoffman et al., 2013) predict network diversity across 15 concepts; network diversity is mean diversity coefficient across nodes. We extracted network measures from the full concept networks and explored how they relate to cognitive measures of conceptual flexibility and stability. Hoffman et al. (2013) use word co-occurrence statistics to quantify the context-dependent variations in word meanings found in language. The authors provide a measure of semantic diversity (SemD) that captures this variability, and we extracted SemD values for our 15 concepts. We also extracted their reported mean cosine similarity of a word's contexts and used this as a measure of semantic stability (which we refer to as SemS). As expected, SemD negatively correlated with SemS across our 15 concepts (r(15)=-0.96, p=<0.0001). The correlations between all measures of interest are shown in Table 2. One of our primary goals was to extract a network measure that reflects conceptual flexibility. We used SemD (Hoffman et al., 2013) as a benchmark for conceptual flexibility and determined whether our hypothesized network measures of flexibility correlated with SemD across our 15 concepts. A priori, we hypothesized that the mean diversity (i.e., the average of a concept network's diversity coefficients across nodes) could reflect conceptual flexibility. This network measure captures the extent to which properties within a concept associate with different modules, or property clusters. Another possible candidate measure was network modularity, which reflects the extent to which a concept's network can be 18 partitioned into separate property clusters. Network modularity was not significantly predicted by either SemD (r(15)=0. 22, p>0.4) or SemS (r(15)=-0.19, p>0.5). On the other hand, mean diversity was positively predicted by SemD (r(15)=0.56, p=0.03; Fig. 4) and negatively predicted by SemS (r(15)=-0.60, p=0.02). Mean diversity was not significantly predicted by mean node strength (r(15)=-0.08, p>0.7), mean node degree (r(15)=0.42, p=0.12), or total number of network edges (r(15)=0.3, p>0.2). These results suggest that the network measure of mean diversity is a strong candidate for a quantitative measure of conceptual flexibility. We also assessed the core-periphery structure for each concept network, which determines how well a network can be divided into a densely connected core and a sparsely connected periphery. If the core of a concept network corresponds to the notion of a context-independent conceptual "core", we would expect that more stable (i.e., less flexible) concepts would have networks with a stronger core-periphery structure. Consistent with this prediction, core fit was positively predicted by SemS (r(15)=0.54, p=0.04), though the relationship with SemD was only marginally significant (r(15)=- 0.50, p=0.06). Furthermore, mean diversity and core fit were negatively correlated (r(15)=-0.61, p=0.02), suggesting that these measures may be used to capture conceptual flexibility and stability, respectively. We also found that classification accuracy using the traditional vector model was positively correlated with core fit (r(15)=0.56, p=0.03). This suggests that standard cognitive models perform better on more stable concepts, highlighting the need for a model that can adequately capture conceptual flexibility. Discussion Here our goal was to model basic-level concepts using graph-theoretical networks within a compositional conceptual framework. We argue that the within-concept statistics encoded in these models capture useful, concept- specific information. Using standard network science tools, we further reveal the usefulness of these models by extracting formal metrics that relate to cognitive notions of conceptual flexibility and stability. Our concept network models capture the particular conceptual properties that are associated with an individual concept along with those properties' 19 concept-specific covariation statistics. Individual concept networks are distinct in that properties relate to each other in different ways across basic- level concepts. For example, BLACK and SOFT may co-vary with each other in BANANA, but BLACK and FIRM may co-vary with each other in CHOCOLATE. We found that our network models could successfully discriminate between new conceptual exemplars, suggesting that these within-concept statistics differ reliably between basic-level concepts. These results emerged out of a classification analysis based on eigendecomposition of our concept networks. Eigendecomposition of graphs has previously been used to assess the correspondences between anatomical brain network structure and patterns of functional activation (Medaglia et al., 2017); here we adapted this method to assess the correspondences between conceptual structure and feature-patterns for individual conceptual exemplars. We found that concept networks can simultaneously encode multi-property relationships (i.e. within-concept statistics) and strong single-property contributions, suggesting ways in which information might be organized within the conceptual system and also establishing an exciting new direction for further investigation. A model structured using within-concept statistics provides a framework in which varied yet appropriate instantiations of a concept may be flexibly activated. An APPLE network may contain a strong connection between CRUNCHY + FRESH and between SOFT + BAKED, enabling the conceptual system to know what sets of properties should be activated in a particular APPLE instance -- for example, in the representations evoked by "apple picking" versus "apple pie." The property-covariation statistics for a given concept will determine which sets of properties tend to be co-activated, and how individual properties relate to those sets and to each other. We thus sought to use our compositional concept network models, which contain within-concept statistics, to extract quantitative measures of these phenomena. We found that mean-diversity and core-periphery structure can be interpreted as measures of conceptual flexibility and stability, respectively: a concept network-model's mean-diversity positively predicts semantic diversity (SemD; Hoffman et al., 2013), a network-model's core- periphery fit positively predicts semantic stability (mean cosine similarity; Hoffman et al., 2013), and these two network measures are negatively related to each other across our concepts. Our results also suggest that traditional property-vector models are better at capturing the representation of stable versus flexible concepts, suggesting that a different kind of conceptual model may be necessary to capture the intrinsic flexibility of the 20 conceptual system. We argue that a network-based model of basic-level concepts is one such option. Network mean-diversity was extracted by averaging over individual nodes' diversity coefficients, a version of the participation coefficient calculated using Shannon entropy. In network neuroscience, participation coefficients have been related to the flexibility of functional network hubs. As discussed above, networks can be partitioned into communities of modules, and nodes can differentially participate in these modules: nodes that only have connections to one module have low participation coefficients, whereas nodes that have connections to many modules have high participation coefficients. This measure is used to differentiate between different kinds of network hubs. High-degree nodes that have low participation coefficients are classified as "provincial" hubs, since they are only connected to one local module. High-degree nodes that have high participation coefficients are classified as "global" or "connector" hubs, since they have connections to many modules across the network (van den Heuvel, 2013; Power et al., 2013). Empirical studies have revealed that the fronto-parietal and cingulo- opercular networks contain high-participation nodes, suggesting that these regions contain functional network hubs (Sporns, 2014). It is further argued that "flexible network hubs" are characterized by their ability to connect with a diverse set of brain regions (Sporns, 2014). Networks have varying degrees of flexibility with respect to how their states can shift with time. On a higher level, networks can differ in how their underlying structure changes with time. In order to study the fluctuations in such dynamic neural networks, Bassett et al. (2011) quantify changes in community structure over time. Instead of extracting a measure from a single network, they define flexibility by analyzing multiple, consecutive networks and determining whether individual nodes either maintain or switch allegiances between communities. Given that networks can exhibit flexibility at these different levels of analysis, we might expect some features of static networks to support flexibility within dynamic networks. We believe that node-participation -- and node-diversity, specifically -- captures a network characteristic of community allegiance similar to Bassett et al. (2011). Furthermore, high-participation or high-diversity nodes (i.e., global hubs) should play a key role in flexibility because their dense, variable connections mediate transitions between network states. We thus argue that diversity might index flexibility across multiple levels of analysis, and that concept network diversity can be reasonably interpreted as a 21 measure of conceptual flexibility. We do not rule out the possibility that there are other candidate network measures that can also capture conceptual flexibility, and intend to explore this moving forward. Other frameworks have the potential to capture the flexibility of the conceptual system; these include attractor networks (e.g., Cree et al., 1999; 2006; Rodd et al., 2004) and recent updates of the hub-and-spoke model (Lambon Ralph et al., 2016). The concept network framework proposed here is not in opposition with these other approaches; the development and implementation of all of these methods will greatly benefit our understanding of the semantic system. However, we do believe that a network science approach to conceptual knowledge has its unique advantages. Most broadly, the vast network science methodological toolkit allows us to translate our analyses between cognitive and neural levels of analysis. Given a sample of concept networks and extracted measures of interest, we can observe the functional neural networks recruited for conceptual processing and explore any correspondences between networks across levels. Network neuroscientists have previously forged links to cognitive processes such as motor-sequence learning (Bassett et al., 2011) and cognitive control (Medaglia et al., 2018), setting a precedent for the application of these methods to cognitive science. The intersection of network science with control theory is another direction that may prove useful for our current purposes. Network controllability refers to the ability to move a network into different network states, and has been applied to structural brain networks in order to shed insights into how the brain may guide itself into easy- and difficult-to-reach functional states (Gu et al., 2015). There have been additional attempts to link brain network controllability to cognitive control (Medaglia, 2018). The application of control theory to concept networks may provide an additional way to quantify conceptual flexibility by identifying nodes that are well-positioned to drive the brain into diverse, specific, or integrated states. Perhaps concept networks that are more controllable overall -- that is, networks in which it is easier to reach varied network states -- correspond to concepts that are more cognitively flexible. The concept network framework permits the application of spreading activation models to assess how information flows through these networks. De Deyne et al. (2016) paired a spreading activation model over relational semantic networks and found that their network models could make accurate 22 predictions regarding semantic similarity judgments. In our case, we could use spreading activation models to observe patterns of network activity as a result of different inputs (i.e., contexts). For example, we can attempt to model how conceptual information differs when presented in an adjective- noun phrase: a vector representing adjectival information could be provided as input to the networks, resulting in a specific pattern of activity across the nodes. More generally, these techniques might provide a way to model context-dependent conceptual meaning. We acknowledge that our proposed concept network model framework has some limitations. First, a large amount of data needs to be collected to enable the calculation of within-concept statistics. However, now that we have established the feasibility and usefulness of these models, it is possible to develop online platforms that can streamline data collection. Second, there is a sense in which some of the methodological decisions are arbitrary: for example, the number of properties to model as network nodes, and the number of concept-states, could influence the resulting networks and extracted measures. These concerns are mitigated, however, by our findings that results calculated from our concept models in Set 1 and Set 2 were comparable, and that the number of concept-states did not significantly predict our network measures of interest. Using standard network filtering methods also reduces these concerns, since it decreases the number of arbitrary parameter decisions required by the experimenter. It is also imperative that the measures extracted from concept networks are interpretable in the context of conceptual knowledge. Only by understanding the potential correspondences between network structure and the structure of conceptual knowledge will these ideas prove useful. Here we have constructed concept network models, confirmed their ability to capture concept-specific information, and extracted network measures that relate to cognitive measures of conceptual flexibility and stability. We believe the application of network science to conceptual knowledge will provide a set of tools that will enable the intrinsic flexibility of the conceptual system to be explored and quantified. 23 Acknowledgements This work was supported by an NSF graduate research fellowship awarded to SHS, DP5-OD021352 awarded to JDM, and National Institute of Health grant R01 DC015359 awarded to STS. Appendix Concept-States (Set 1) CHOCOLATE: bittersweet chocolate, caramel chocolate, chocolate bar, chocolate chips, chocolate syrup, cocoa powder, dark chocolate, chocolate fudge, melted chocolate, milk chocolate, nut chocolate, salted chocolate, white chocolate, hot chocolate BANANA: banana chips, banana pudding, Cavendish banana, fried banana, frozen banana, mashed banana, over-ripe banana, peeled banana, plantain, raw banana, red banana, rotten banana, sliced banana, unripe banana, ripe banana BOTTLE: baby bottle, beer bottle, broken bottle, juice bottle, liquor bottle, medicine bottle, milk bottle, soda bottle, spray bottle, water bottle, wine bottle TABLE: bedside table, changing table, coffee table, conference table, dining table, drafting table, end table, folding table, kitchen table, play table, poker table, pool table, side table, workbench PAPER: butcher paper, cardboard, cardstock, construction paper, envelope, graph paper, legal paper, newspaper, notebook paper, paper towel, papyrus, poster board, printer paper, sandpaper, scrap paper, sketch paper, stationery, tissue paper, toilet paper, wrapping paper, writing paper Concept-States (Set 2) KEY: car key, key to a city, door key, encryption key, garage key, key to my heart, house key, key card, keyboard key, map key, master key, motorcycle key, office key, padlock key, password, piano key, key to a safe, skeleton key PUMPKIN: pumpkin bar, pumpkin bread, pumpkin candle, canned pumpkin, pumpkin cookie, pumpkin in a field, Halloween pumpkin, Jack-O-Lantern, pumpkin latte, pumpkin muffin, pumpkin pie, pumpkin puree, rotten pumpkin, pumpkin seeds, smashed pumpkin, pumpkin soup, pumpkin spice, whole pumpkin, Thanksgiving pumpkin 24 GRASS: astroturf, bamboo, barley grass, bent grass, grass clippings, crab grass, dead grass, hay, lawn grass, lemongrass, marijuana, oat grass, overgrown grass, grass seeds, sod, wheatgrass COOKIE: almond cookie, butter cookie, chocolate cookie, chocolate chip cookie, Christmas cookie, cookie cake, cookie dough, ginger snap, Girl Scout cookie, lemon cookie, M&M cookie, macadamia nut cookie, macaroon, mint cookie, no-bake cookie, oatmeal raisin cookie, Oreo cookie, peanut butter cookie, shortbread, snickerdoodle, sugar cookie, wafer cookie PICKLE: bread and butter pickles, canned pickles, pickle chips, chopped pickles, cucumber pickles, dill pickles, garlic pickles, gherkins, hamburger pickles, homemade pickles, jarred pickles, kosher pickles, relish, sliced pickles, sandwich pickles, pickle spears, whole pickles PILLOW: airplane pillow, bed pillow, body pillow, cotton pillow, couch pillow, decorative pillow, down pillow, feather pillow, foam pillow, hypo-allergenic pillow, memory foam pillow, neck pillow, silk pillow, throw pillow, travel pillow KNIFE: bread knife, butcher knife, butter knife, cheese knife, chef's knife, dagger, hunting knife, jackknife, machete, paring knife, pocket knife, steak knife, switchblade, sword, throwing knife, utility knife WOOD: wood blocks, wood chips, chopped wood, wood fence, firewood, floor, wood furniture, log, lumber, wood paneling, paper, planks, plywood, wood pulp, sticks, tree, cedar wood, cherry wood, maple wood, oak wood, pine wood, walnut wood PHONE: android phone, antique phone, broken phone, car phone, cell phone, emergency phone, flip phone, home phone, iPhone, land line, pay phone, rotary phone, satellite phone, smart phone, wall phone, wireless phone CAR: broken down car, compact car, convertible, coupe, electric car, family car, hatchback, hybrid car, jeep, luxury car, pickup truck, race car, rental car, sedan, sports car, station wagon, SUV, toy car, truck, used car, van References Amaral, L. A. N., Scala, A., Barthelemy, M., & Stanley, H. E. (2000). Classes of small- world networks. Proceedings of the national academy of sciences, 97(21), 11149- 11152. Barabási, A. L. (2016). Network science. Cambridge university press. 25 concepts. Memory & Cognition, 10(1), 82-93. Barsalou, L. W. (1982). Context-independent and context-dependent information in Bassett, D. S., & Bullmore, E. T. (2017). Small-world brain networks revisited. The Bassett, D. S., Wymbs, N. F., Porter, M. A., Mucha, P. J., Carlson, J. M., & Grafton, S. Neuroscientist, 23(5), 499-516. T. (2011). Dynamic reconfiguration of human brain networks during learning. Proceedings of the National Academy of Sciences, 108(18), 7641-7646. Bassett, D. S., Wymbs, N. F., Rombach, M. P., Porter, M. A., Mucha, P. J., & Grafton, S. T. (2013). Task-based core-periphery organization of human brain dynamics. PLoS computational biology, 9(9), e1003171. networks, 21(4), 375-395. Borgatti, S. P., & Everett, M. G. (2000). Models of core/periphery structures. Social Collins, A. M., & Loftus, E. F. (1975). A spreading-activation theory of semantic Collins, A. M., & Quillian, M. R. (1969). Retrieval time from semantic memory. Journal Cree, G. S., McRae, K., & McNorgan, C. (1999). An attractor model of lexical of verbal learning and verbal behavior, 8(2), 240-247. processing. Psychological review, 82(6), 407. conceptual processing: Simulating semantic priming. Cognitive Science, 23(3), 371- 414. Cree, G. S., McNorgan, C., & McRae, K. (2006). Distinctive features hold a privileged status in the computation of word meaning: Implications for theories of semantic memory. Journal of Experimental Psychology: Learning, Memory, and Cognition, 32(4), 643. De Deyne, S., Navarro, D. J., Perfors, A., & Storms, G. (2016). Structure at every scale: A semantic network account of the similarities between unrelated concepts. Journal of Experimental Psychology: General, 145(9), 1228. Griffiths, T. L., Steyvers, M., & Tenenbaum, J. B. (2007). Topics in semantic Gu, S., Pasqualetti, F., Cieslak, M., Telesford, Q. K., Alfred, B. Y., Kahn, A. E., ... & representation. Psychological review, 114(2), 211. Bassett, D. S. (2015). Controllability of structural brain networks. Nature communications, 6, 8414. Hoffman, P., Ralph, M. A. L., & Rogers, T. T. (2013). Semantic diversity: A measure of semantic ambiguity based on variability in the contextual usage of words. Behavior research methods, 45(3), 718-730. 26 Huang, W., Bolton, T.A.W., Medaglia, J.D., Bassett, D.S., Ribeiro, A., & Van De Ville, D. (In Press). A Graph Signal Processing Perspective on Functional Brain Imaging. IEEE. of the Royal Society of London B: Biological Sciences, 268(1482), 2261-2265. i Cancho, R. F., & Solé, R. V. (2001). The small world of human language. Proceedings Kenett, Y. N., Anaki, D., & Faust, M. (2014). Investigating the structure of semantic Landauer, T. K., & Dumais, S. T. (1997). A solution to Plato's problem: The latent networks in low and high creative persons. Frontiers in Human Neuroscience, 8, 407. semantic analysis theory of acquisition, induction, and representation of knowledge. Psychological review, 104(2), 211. Massara, G. P., Di Matteo, T., & Aste, T. (2016). Network filtering for big data: triangulated maximally filtered graph. Journal of complex Networks, 5(2), 161-178. McRae, K. (2004). Semantic memory: Some insights from feature-based connectionist attractor networks. The psychology of learning and motivation: Advances in research and theory, 45, 41-86. McRae, K., Cree, G. S., Westmacott, R., & Sa, V. R. D. (1999). Further evidence for feature correlations in semantic memory. Canadian Journal of Experimental Psychology/Revue canadienne de psychologie expérimentale, 53(4), 360. McRae, K., De Sa, V. R., & Seidenberg, M. S. (1997). On the nature and scope of featural representations of word meaning. Journal of Experimental Psychology: General, 126(2), 99. Medaglia, J. D., Huang, W., Karuza, E. A., Kelkar, A., Thompson-Schill, S. L., Ribeiro, A., & Bassett, D. S. (2017). Functional alignment with anatomical networks is associated with cognitive flexibility. Nature Human Behaviour, 1. arXiv:1801.08806. Evaluation, 41(2), 209-214. Medaglia (2018). Clarifying cognitive control and the controllable connectome. Miller, G. A., & Fellbaum, C. (2007). WordNet then and now. Language Resources and Power, J. D., Schlaggar, B. L., Lessov-Schlaggar, C. N., & Petersen, S. E. (2013). Evidence for hubs in human functional brain networks. Neuron, 79(4), 798-813. Ralph, M. A. L., Jefferies, E., Patterson, K., & Rogers, T. T. (2017). The neural and computational bases of semantic cognition. Nature Reviews Neuroscience, 18(1), 42. 27 Randall, B., Moss, H. E., Rodd, J. M., Greer, M., & Tyler, L. K. (2004). Distinctiveness and correlation in conceptual structure: Behavioral and computational studies. Journal of Experimental Psychology: Learning, Memory, and Cognition, 30(2), 393. semantic ambiguity in word recognition. Cognitive Science, 28(1), 89-104. Rodd, J. M., Gaskell, M. G., & Marslen-Wilson, W. D. (2004). Modelling the effects of Rubinov, M., & Sporns, O. (2010). Complex network measures of brain connectivity: Rubinov, M., & Sporns, O. (2011). Weight-conserving characterization of complex Rubinov, M., Ypma, R. J., Watson, C., & Bullmore, E. T. (2015). Wiring cost and functional brain networks. Neuroimage, 56(4), 2068-2079. uses and interpretations. Neuroimage, 52(3), 1059-1069. topological participation of the mouse brain connectome. Proceedings of the National Academy of Sciences, 112(32), 10032-10037. memory: A featural model for semantic decisions. Psychological review, 81(3), 214. Smith, E. E., Shoben, E. J., & Rips, L. J. (1974). Structure and process in semantic Sporns, O. (2014). Contributions and challenges for network models in cognitive Steyvers, M., & Tenenbaum, J. B. (2005). The Large-scale structure of semantic neuroscience. Nature neuroscience, 17(5), 652. networks: Statistical analyses and a model of semantic growth. Cognitive science, 29(1), 41-78. Tumminello, M., Aste, T., Di Matteo, T., & Mantegna, R. N. (2005). A tool for filtering information in complex systems. Proceedings of the National Academy of Sciences of the United States of America, 102(30), 10421-10426. knowledge. Trends in cognitive sciences, 5(6), 244-252. Tyler, L. K., & Moss, H. E. (2001). Towards a distributed account of conceptual van den Heuvel, M. P., & Sporns, O. (2013). Network hubs in the human brain. Trends in Van Rensbergen, B., Storms, G., & De Deyne, S. (2015). Examining assortativity in the cognitive sciences, 17(12), 683-696. mental lexicon: Evidence from word associations. Psychonomic bulletin & review, 22(6), 1717.
1905.02405
1
1905
2019-05-07T08:37:31
Neuroplasticity in adult human visual cortex
[ "q-bio.NC" ]
Between 1 to 5 out of 100 people worldwide has never experienced normotypic vision due to a condition called amblyopia, and about 1 out of 4000 suffer from inherited retinal dystrophies that progressively lead them to blindness. While a wide range of technologies and therapies are being developed to restore vision, a fundamental question still remains unanswered: would the adult visual brain retain a sufficient plastic potential to learn how to see after a prolonged period of abnormal visual experience? In this review we summarize studies showing that the visual brain of sighted adults retains a type of developmental plasticity, called homeostatic plasticity, and this property has been recently exploited successfully for adult amblyopia recover. Next, we discuss how the brain circuits reorganizes when visual stimulation is partially restored by means of a bionic eye in late blinds with Retinitis Pigmentosa. The primary visual cortex in these patients slowly became activated by the artificial visual stimulation, indicating that sight restoration therapies can rely on a considerable degree of spared plasticity in adulthood.
q-bio.NC
q-bio
Neuroplasticity in adult human visual cortex Elisa Castaldi1, Claudia Lunghi2 and Maria Concetta Morrone3,4 1 Department of Neuroscience, Psychology, Pharmacology and Child health, University of Florence, Florence, Italy 2 Laboratoire des systèmes perceptifs, Département d'études cognitives, École normale supérieure, PSL University, CNRS, 75005 Paris, France 3 Department of Translational Research and New technologies in Medicine and Surgery, University of Pisa, Pisa, Italy 4 IRCCS Stella Maris, Calambrone (Pisa), Italy Abstract Between 1-5:100 people worldwide has never experienced normotypic vision due to a condition called amblyopia, and about 1:4000 suffer from inherited retinal dystrophies that progressively lead them to blindness. While a wide range of technologies and therapies are being developed to restore vision, a fundamental question still remains unanswered: would the adult visual brain retain a sufficient plastic potential to learn how to 'see' after a prolonged period of abnormal visual experience? In this review we summarize studies showing that the visual brain of sighted adults retains a type of developmental plasticity, called homeostatic plasticity, and this property has been recently exploited successfully for adult amblyopia recover. Next, we discuss how the brain circuits reorganizes when visual stimulation is partially restored by means of a 'bionic eye' in late blinds with Retinitis Pigmentosa. The primary visual cortex in these patients slowly became activated by the artificial visual stimulation, indicating that sight restoration therapies can rely on a considerable degree of spared plasticity in adulthood. Keywords Keywords: 7T fMRI; amblyopia; binocular rivalry; bionic eye; blindness; cortical excitability; critical period; cross-modal plasticity; retinal prosthesis; retinitis pigmentosa; short-term monocular deprivation; visual restoration Highlights  Short-term monocular deprivation triggers homeostatic plasticity in adult humans  Physical activity and inverse occlusion promote visual recovery in amblyopic adults  Intra-modal and cross-modal brain reorganization occur in late-blind individuals  Sight restoration in late-blinds promotes early visual cortex re-organization  The adult visual system retains residual plasticity well after the critical period Index 1. Introduction 2. Cortical plasticity in sighted adults revealed by short-term visual deprivation 2.1. Behavioral proxies for plasticity 2.2. Short-term visual deprivation alters visual neural responses 2.3. Neurochemical changes following short-term visual deprivation 2.4. Therapy for amblyopia 3. Cortical plasticity after vision loss following ophthalmological diseases 3.1. Retinotopic remapping of the visual cortex 3.2. Cross-modal plasticity in blind individuals 3.3. Cortical plasticity and sight restoration 4. Conclusions and further directions 1. Introduction Neuroplasticity is the ability of the nervous system to adapt and optimize its limited resources in response to physiological changes, injuries , new environmental demands and sensory experiences (Pascual-Leone et al., 2005). Early in the individual's development, during the so called critical period, the intrinsic plastic potential of the nervous systems is maximal and sensory deprivation events can cause profound morphological alterations of sensory cortices preventing the normal development of sensory functions (Berardi et al., 2000; Wiesel and Hubel, 1963). Classical studies in the visual system of kittens and non-human primates demonstrated that even few days of abnormal visual experience during the period of high susceptibility can cause severe visual impairment that cannot be recovered even after prolonged periods of restored vision (Hubel and Wiesel, 1970; Hubel et al., 1977; Wiesel and Hubel, 1963, 1965b). Visual cortical plasticity is typically assessed by using the experimental paradigm of monocular deprivation: occluding one eye few weeks after birth causes atrophy of the lateral geniculate nucleus (LGN) layers representing the deprived eye, reorganization of primary visual cortex (V1) ocular dominance columns, with ocular dominance shifting in favor of the open eye and with only a few neurons responding to the deprived eye (Hubel and Wiesel, 1970; Hubel et al., 1977; Wiesel and Hubel, 1963). At the functional level, this shift of ocular dominance is reflected in lower visual acuity and reduced response to stimulation of the deprived eye, a condition known as amblyopia (for a review see: Levi and Carkeet, 1993). In adults, after the closure of the critical period, amblyopia cannot be induced or treated: in adult cats monocular deprivation has only minor or no effects, suggesting that the visual cortex retains very little or no experience-dependent plasticity in adulthood (Wiesel and Hubel, 1963). Similarly, in humans early visual deprivation or suboptimal visual experience (for example due to untreated congenital cataracts, astigmatism, myopia, etc.) leads to plastic reorganization of the visual cortices and permanent visual impairments that can hardly be recovered in adulthood even after corrections or visual restoration (Braddick and Atkinson, 2011; Fine et al., 2002, 2003; Maurer et al., 2005, 2007). The ocular dominance imbalance induced by early altered monocular inputs can be efficiently treated by occluding the dominant eye for prolonged periods of time (occlusion therapy) but only within the critical period (Webber and Wood, 2005), whereas only modest improvements of visual function are observed when therapy is performed in adulthood (Fronius et al., 2014). While neuroplasticity is preserved in adulthood for higher-level functions, endorsing learning and memory (Fuchs and Flugge, 2014), the absence of experience-dependent changes observed after the closure of the critical period led researchers to consider the visual system, and in particular early visual cortex, as hard-wired and with no spared plastic potential, especially for ocular dominance. However, several recent lines of evidence have put this assumption into question: in animal models, ocular dominance plasticity can be reactivated after the closure of the critical period by manipulating the visual cortex excitability, either pharmacologically or through environmental enrichment and physical activity (Baroncelli et al., 2010, 2012; Berardi et al., 2015; Harauzov et al., 2010; He et al., 2006; Hensch and Fagiolini, 2005; Hensch et al., 1998; Maya Vetencourt et al., 2008; Pizzorusso et al., 2002; Sale et al., 2007, 2014; Spolidoro et al., 2009). In adult humans, evidence of preserved visual plasticity has been demonstrated by behavioral and neural changes associated with perceptual learning (Beyeler et al., 2017; Dosher and Lu, 2017; Fiorentini and Berardi, 1980; Watanabe and Sasaki, 2015), short term visual deprivation (Binda et al., 2018; Binda and Lunghi, 2017; Lunghi et al., 2011, 2013, 2015a,b, 2018; Lunghi and Sale, 2015; Zhou et al., 2013, 2014), progressive blinding pathologies and visual restoration therapies (Aguirre et al., 2016; Baseler et al., 2011a; Burton, 2003; Castaldi et al., 2016; Cunningham et al., 2015a,b; Dormal et al., 2015; Heimler et al., 2014). feedback cycles might The cellular mechanisms promoting neuroplasticity during development have been extensively studied in animals (Smith et al., 2009). Cortical wiring in developing neural systems is determined both by molecular cues, responsible for neural migration and formation of synaptic contacts, as well as by activity-dependent mechanisms that fine tune and optimize the number and strength of synaptic connections through Hebbian plasticity (Levelt and Hubener, 2012). This is mediated by Long Term Potentiation and Depression (LTP/LTD), mechanisms that have a crucial role also in learning and memory (Malenka and Bear, 2004) and in defining the critical period (Berardi et al., 2000; Hensch and Quinlan, 2018; Levelt and Hubener, 2012). In addition Hebbian plasticity, another form of plasticity (named homeostatic plasticity) is acting to maintain the overall balance of the network excitability (Maffei and Turrigiano, 2008a). This may have an important role given that LTP/LTD themselves might lead to destabilization of neural circuits: continuous the progressive strengthening/weakening of synaptic connections and to its consequent excessive excitability/loss (Turrigiano and Nelson, 2000). Through mechanisms such as synaptic scaling, synaptic redistribution and changes in neural excitability, the homeostatic plasticity promotes stability in the neural circuits by readjusting the overall level of network activity to optimize responses to sensory experiences and perturbations (Abbott and Nelson, 2000; Maffei and Turrigiano, 2008a,b; Turrigiano, 2012; Turrigiano and Nelson, 2000). After short-term deprivation and during the critical period homeostatic plasticity is boosting the signals from the deprived eye, in the attempt to contrast the deprivation effects. Recent evidence suggests that this homeostatic synaptic rescaling may remain active even after the closure of the critical period and support neuroplasticity throughout the life span. Interestingly, the visual cortex maintains homeostasis by normalizing the individual neural responses with the overall activity of a pool of neurons. This mechanisms is especially important to efficiently control contrast gain (Carandini and Heeger, 2011; Heeger, 1992). It is possible that homeostatic plasticity in adult humans may be implemented using similar cellular mechanisms mediating contrast gain (see later Lunghi et al., 2011) lead to Here, we review evidence for adult plasticity and we speculate on the most likely mechanisms and principles supporting the residual ocular dominance plasticity in adulthood. Deeply understanding the mechanisms regulating adult visual plasticity is crucial for visual restoration in late-blind individuals, considering that the various intervention (retinal prostheses but also pharmacological manipulations) are implemented monocularly: in principle the restored monocular signals might be gated at cortical level if ocular dominance plasticity cannot be endorsed. We start reviewing recent studies reporting perceptual and neural changes induced in healthy sighted individuals by temporary altering their visual experience and how these findings have recently led to the development of a non-invasive training for adult amblyopic patients. We discuss how these phenomena likely reflect homeostatic plasticity in adulthood. In the second part we discuss the neural changes following blinding diseases and in particular the plastic retinotopic remapping of the residual visual input and the cross-modal reorganization of the visual brain when the visual input is altered or interrupted. Finally, we describe the success and falls of the very first attempts of restoring vision in adult blind patients. 2. Cortical plasticity in sighted adults revealed by short- term visual deprivation 2.1 Behavioral proxies for plasticity Recent studies have introduced new behavioral techniques to infer and estimate the degree of residual plasticity for ocular dominance in adult sighted subjects, by combining binocular rivalry (Lunghi et al., 2011) and, more recently, pupillometry (Binda and Lunghi, 2017) with short periods of monocular deprivation. Binocular rivalry is a form of bistable perception that is generated whenever two incompatible images are separately projected to the eyes. In such condition the visual perception alternates between the two monocular images which take turn in dominating visual awareness (Blake and Logothetis, 2002; Levelt, 1965). Binocular rivalry is one of the most robust psychophysical methods used to assess sensory eye dominance (Ooi and He, 2001): under normal conditions, the average time in which the image presented to each eye dominates the observer's perception is similar for the two eyes, reflecting balanced ocular dominance. Lunghi et al. (2011) first observed that visually depriving healthy adult individuals though monocular occlusion with a translucent patch for 150 minutes profoundly altered ocular dominance measured with binocular rivalry. Surprisingly the stimulus presented to deprived eye dominates twice as long as the one displayed in the non-deprived eye after deprivation. The effect, although progressively attenuated, lasted up to 90 or 180 minutes after patch removal (Lunghi et al., 2013). After patch removal the apparent contrast increased for a short time, suggesting an up-regulation of the contrast gain-control mechanisms in the occluded eye that boosted the neuronal responses to compensate for the reduced incoming signal. The altered dynamics of binocular rivalry and the contrast gain enhancement after monocular deprivation suggested the existence of a spared plastic potential in adult individuals: these phenomena most likely reflect a form of homeostatic plasticity attempting to maintain the overall network activity stable and to optimize the individual's new visual experience. Even though the perceptual effects of short-term monocular deprivation might be in principle interpreted as contrast adaptation, growing evidence indicates that they reflect a genuine form of plasticity. For example, even though deprivation alters apparent contrast, the 36% boost in apparent contrast is not sufficient to explain the change in ocular dominance (Lunghi et al., 2011), and when chromatic vision is specifically targeted (using iso-luminant visual stimuli), the effect can out-last the duration of deprivation, lasting for up to 3h (Lunghi et al., 2013). Moreover, two recent studies (Bai et al., 2017; Ramamurthy and Blaser, 2018) have shown that changes in ocular dominance can be observed by altering the monocular input without reducing monocular contrast, pointing to a crucial role of inter-ocular correlation in mediating ocular dominance plasticity. This is consistent with the etiology of amblyopia which can be produced by strabismus (Kiorpes et al., 1998) where both eye receive good vision but with no spatial congruency. Another direct proof that the transient changes in ocular dominance are an expression of homeostatic plasticity is given by the long term improvement of vision in adult amblyopic patients that received a short-term monocular deprivation of the amblyopic eye for 6 short sessions. The improvement in visual acuity of about 2 lines lasted for up to one year, indicating long-term neuroplastic changes (Lunghi et al., 2018). Binda & Lunghi (2017) more recently tackled another biomarker of neuroplasticity in adult humans by demonstrating that monocular deprivation affects spontaneous low frequency oscillations of the pupil diameter at rest, a phenomenon called hippus (Diamond, 2001). The authors measured pupillary oscillations before and after monocular occlusion, following the same paradigm as Lunghi et al. (2011), and found that hippus amplitude increased after visual deprivation and that participants with more pronounced pupillary fluctuations also showed stronger ocular dominance changes in binocular rivalry dynamics. The procedures to measure these two proxies of neuroplasticity are completely non-invasive and suitable for application in clinical populations. 2.2 Short-term visual deprivation alters visual neural responses A substantial number of studies investigating the cortical effects of visual deprivation in adult humans suggested that plasticity might be mediated by changes in the excitation/inhibition balance of the visual cortex (Binda et al., 2018; Boroojerdi et al., 2000; Fierro et al., 2005; Lou et al., 2011; Lunghi et al., 2015a; Pitskel et al., 2007). The first studies (Boroojerdi et al., 2000; Fierro et al., 2005; Pitskel et al., 2007) applied TMS pulses to the occipital cortex sufficient to elicit light perception (phosphene) in absence of visual stimulation. The minimum intensity needed to elicit phosphene perception (PT, phosphene threshold) is an indirect measure of cortical excitability. Boroojeredi et al. (2000) showed that after 45 minutes of binocular light deprivation PT was reduced in healthy adult subjects, suggesting increased cortical excitability and this effect persisted for the entire deprivation period (180 minutes). Re- exposure to light reverted the process and PT returned to pre-deprivation values within 120 minutes. Interestingly the same study provided also another measure of cortical responsiveness, independent of the subjects' perception: neural activity in striate and extrastriate cortices as measured by BOLD responses with functional magnetic imaging (fMRI) was enhanced after 60 minutes of blindfolding and the increased fMRI signal persisted for at least 30 minutes after re-exposure to light. Monocular deprivation might cause different cortical effects with respect to blindfolding, possibly due to inter-ocular competition which is absent in case of binocular deprivation. In fact, evidence from animal studies has shown that, compared with monocular deprivation, binocular deprivation performed during the critical period induces a more modest reorganization of visual cortical circuits (Wiesel and Hubel, 1965a). In adult humans, monocular deprivation was shown to cause a decrease in cortical excitability, rather than an increase, as measured by TMS PT (Lou et al., 2011). However a recent study measuring pattern-onset visual evoked potentials (VEPs) before and after monocular deprivation described a more complex pattern of results with opposite effects for the two eyes (Lunghi et al., 2015a). The amplitude of the C1 component of the VEP responses, typically reflecting the earliest stage of visual processing in V1 (Di Russo et al., 2002), and the peak in alpha band after monocular deprivation were enhanced for the deprived eye, but reduced for the non-deprived eye. Moreover the amplitude of later components, such as P1 and P2, were equally altered by monocular deprivation, suggesting that the variations in cortical excitability propagate to extrastriate areas and modulate feed-back projection to V1. Overall these results, in line with the perceptual changes observed after short-term monocular deprivation (described in section 2.1), suggest that the deprived and non-deprived eye are respectively strengthened or weakened after monocular deprivation, reflecting antagonistic homeostatic short-term plasticity in the two eyes. Recently Binda et al. (2018) exploited the enhanced resolution and signal to noise ratio provided by ultra-high field (7T) fMRI, to track ocular driven changes of BOLD responses in V1 before and after 2h of monocular deprivation. During the scanning participants' eyes were separately stimulated with either high contrast low- and high- band-pass noise (optimized to differentially stimulate the magno- and parvocellular pathways respectively) or with a luminance matched uniform background (Fig.1A). BOLD responses to the high spatial frequency stimuli were strongly affected by monocular deprivation, in opposite directions for the two eyes: the percentage BOLD signal measured in V1 elicited by stimulation of the deprived eye and non-deprived eye was respectively enhanced and reduced with respect to pre-deprivation values (Fig 1B). The authors further calculated for each voxel an index of Ocular Dominance defined as the response difference to the deprived and non-deprived eye, which reflects the eye preference of a given particular voxel (average biased signal), although not strictly coinciding with the ocular dominance columns. Before deprivation, the ocular dominance indices were symmetrically distributed around zero, reflecting balanced V1 activity elicited by the two eyes. However, after deprivation the index strongly shifted toward a stronger activation from the deprived eye. Interestingly the voxels originally preferring the deprived eye did not change their preference (i.e. the average signal measured in these voxels continued to be biased toward the same eye), whereas the voxels originally preferring the non-deprived eye swapped their preference and were most strongly activated by stimulation of the deprived eye. Importantly the deprivation effect in BOLD responses correlated with the perceptual effect of deprivation as measured by increased mean phase duration for the deprived eye during binocular rivalry performed outside the scanner with the classic paradigm (Fig. 1 C, D). Spatial frequency selectivity in V1 was also affected by deprivation: only responses elicited by high-spatial frequencies stimulation of the deprived eye were significantly reduced, while responses to low-spatial frequencies stimulation of the same eye as well as to the whole frequencies range of the non-deprived eye were unaffected. Population Spatial Frequency Tuning of V1 confirmed these effects and the shift of selectivity toward higher spatial frequency in the deprived eye correlated with the phase duration during binocular rivalry. Interestingly, the increase in BOLD response to the high-spatial frequency stimulation of the deprived eye was strongest in V1, V2, V3, attenuated but still significant in V4, while it was absent in V3a and hMT+. The enhanced signals measured along the ventral but not along the dorsal pathway suggested that plasticity acted more strongly on the parvocellular rather than magnocellular pathway. This interpretation is also supported by the evidence that deprivation-induced changes in binocular rivalry dynamics with chromatic equi- luminant gratings resulted in much longer-lasting effects with respect to those measured with luminance-modulated gratings (Lunghi et al., 2013). In sum, these experiments showed that complex patterns of neuroplastic responses can be induced in the visual cortex of sighted adults by short periods of visual deprivation, which rapidly alter the visual cortex's excitability and responsiveness to homeostatically adapt to the altered visual experience. 2.3 Neurochemical changes following short-term visual deprivation The molecular mechanisms underlying experience-dependent plasticity have been widely investigated in animals (Berardi et al., 2003; Heimel et al., 2011). The maturation of intracortical inhibition has been proved to play a crucial role in regulating the progression of critical period for ocular dominance and visual acuity (Berardi et al., 2003; Fagiolini et al., 2004; Fagiolini and Hensch, 2000; Hensch et al., 1998; Huang et al., 1999; Speed et al., 1991). For instance, increasing intracortical inhibition was shown to anticipate the opening and closure of the critical period for monocular deprivation in mice (Fagiolini and Hensch, 2000; Hanover et al., 1999; Huang et al., 1999), and transgenic animals lacking a GABA-synthesizing enzyme showed deficient ocular dominance plasticity to monocular deprivation which could be restored by increasing inhibitory transmission with benzodiazepines (Hensch et al., 1998). In adulthood, the increased inhibition may be a limiting factor for cortical plasticity and reducing GABAergic inhibition was shown to partially restore ocular dominance plasticity in adult rats and promote recovery from amblyopia (Harauzov et al., 2010; Maya Vetencourt et al., 2008). Interestingly, environmental enrichment and physical activity were recently found to be associated with reduced inhibitory tone in the rat's visual cortex, providing a potential non-invasive strategy to promote recovery from amblyopia (Baroncelli et al., 2010, 2012; Sale et al., 2007; Stryker, 2014). In line with these results, one study recently showed that GABAergic inhibition plays a key role in promoting ocular dominance plasticity also in adult humans (Lunghi et al., 2015b). Participants underwent one psychophysical test measuring ocular dominance by means of binocular rivalry and one 7T MR spectroscopy session before and after 150 min of monocular occlusion (same procedure as described in section 2.1). The perceptual changes triggered by deprivation (resulting in the deprived eye dominating over the non-deprived eye) were associated with decreased GABA concentration in V1, at least when GABA concentration was assessed at rest, while participants kept their eyes closed. Specifically, participants with greater decrease in resting GABA concentration showed the greatest perceptual effects, i.e. perceptual boost and dominance of the deprived eye. Interestingly, other studies showed that fMRI responses in visual cortex inversely correlated with GABA concentration, potentially suggesting a link between GABA level and cortical excitation-inhibition balance (Donahue et al., 2010; Muthukumaraswamy et al., 2009). Of course GABAergic circuits are not the only ones mediating plasticity in visual cortex (Berardi et al., 2003). In animals, ocular dominance plasticity can be restored also by enhancing excitatory neurotransmission systems such as serotoninergic (Maya Vetencourt et al., 2008) and cholinergic (Morishita et al., 2010) systems. In humans Boroojeredi et al. (2001) applied TMS over occipital cortex and measured phosphene thresholds in adult sighted participants under the effect of various drugs interfering with synaptic plasticity before and after a period of light deprivation. They found that the rapid plastic changes typically triggered by light deprivation (here quantified as phosphene detection thresholds) were blocked when participants were under the effect of lorazepam (which enhance the functioning of GABAa receptors), dextromethorphan (NMDA receptors antagonist) and scopolamine (a muscarinic receptor antagonist), thus pointing at a key role of GABA, NMDA and cholinergic receptors in mediating rapid plastic changes after visual deprivation. Finally, in light of the recent results by Binda & Lunghi (2017) showing both increased pupillary hippus at rest and enhanced eye-dominance of the deprived eye during binocular rivalry after monocular deprivation, the role of the neurotransmitter norepinephrine (NE) in mediating homeostatic plasticity should be further investigated in adult humans. This neurotransmitter may indeed constitute the common source of visual cortical excitability underlying these phenomena, given its known role in regulating both pupil diameter modulation (Joshi et al., 2016) and visual cortical plasticity (Kasamatsu et al., 1979, 1981). Overall, these studies point at a spared plastic potential in the adult visual cortex beyond the critical period which can be reactivated by altering the excitability level of visual cortex either by pharmacologically targeting several neuromodulator systems or by manipulating sensory and motor experience, such as being exposed to abnormal visual experiences even for a short time period, or being exposed to an enriched environment and physical activity. 2.4 Therapy for amblyopia Inspired by studies on animal models showing that physical exercise triggers visual cortical plasticity, modulates visual cortex excitability and increases neurotrophic factors (Baroncelli et al., 2010, 2012; Sale et al., 2007), recent studies tested whether similar effects could be obtained in adult humans (Lunghi and Sale, 2015; Lunghi et al., 2018). Lunghi & Sale (2015) tested binocular rivalry dynamics in sighted participants before and after monocular deprivation (still with the same paradigm described in 2.1) while varying the level of physical activity performed by participants during the deprivation period. In the 'inactive condition' participants were required to watch a movie while sitting on a chair, while in the 'physical activity' condition they watched a movie while intermittently cycling on a stationary bike. With respect to the control inactive condition, the perceptual effect of deprivation on binocular rivalry dynamic was much stronger when participant performed physical activity throughout the 2h period tested after eye-patch removal, suggesting that physical activity had further boosted homeostatic plasticity. The beneficial effect of moderate physical activity for triggering neuroplasticity was recently combined with short-term inverse occlusion to promote the visual recovery in adult anisometropic amblyopes (Lunghi et al., 2018). Six 2h long training sessions over a 4 weeks period consisting in simultaneous physical activity (intermitted cycling) and occlusion of the amblyopic eye restored visual acuity in all patients and stereopsis in six of them with improvement lasting up to 1 year. Although these results should be replicated in a larger sample, this non-invasive training paradigm seemed to successfully boost visual plasticity in adulthood and to constitute a valid approach to treat amblyopia. 3. Cortical plasticity after vision loss following ophthalmological diseases 3.1 Retinotopic remapping of the visual cortex The blind brain certainly constitute a unique opportunity for studying plasticity in adulthood. A highly studied phenomena, sometimes considered as an index of neuroplasticity, is the ability of the visual cortex to remap the retinotopic organization of the neuronal receptive fields following retinal lesions (for a recent review see: Dumoulin and Knapen, 2018). The first evidence for this phenomenon was reported in adult cats and monkeys (Gilbert and Wiesel, 1992): after retinal lesion V1 receptive fields near the border of the lesion projection zone (LPZ) underwent an immediate enlargement, and two months after the regions silenced by the lesion were found to represent loci surrounding the scotoma. The receptive fields shift leading to complete filling in of the scotoma was attributed to long-range horizontal connections within V1 rather than to spared geniculate afferent connections: none of the changes observed in the recovered cortex where observed in the lateral geniculate nucleus, perhaps due to its reduced plasticity with respect to higher level cortical regions (however see section 3.3). More recently these findings have been strongly questioned by a combined neurophysiological and fMRI study that failed to record normal responsivity in adult macaque V1 during 7.5 months of follow-up after retinal lesion (Smirnakis et al., 2005). The authors found no change in the BOLD-defined LPZ border and suggested that cortical reorganization is not needed to explain the apparent size of the LPZ (Smirnakis et al., 2005). Incongruences with respect to previous results were attributed to several factors, including sampling biases and differences in the recording methods (one recording sub-sets of single neurons, the other reflecting average activation of ensembles of cells) and originated an intense debate (for details on this debeat see: Calford et al., 2005; Sereno, 2005; Smirnakis et al., 2005; Wandell and Smirnakis, 2009). The ability of the adult visual cortex to reorganize has also been studied in retinal dystrophies, namely macular degeneration (MD, both age-related MD and juvenile MD) and retinitis pigmentosa (RP). Both diseases create scotomas in patients' central (MD) and peripheral (RP) visual field that expand with illness progression. Some fMRI studies described large-scale cortical reorganization of visual processing in response to retinal disease by showing that the regions of V1 matching the patients' scotoma are remapped to respond to stimuli outside it both in MD (Baker et al., 2005, 2008; Dilks et al., 2009, 2014; Schumacher et al., 2008) and in RP patients (Ferreira et al., 2017). However, others have found the LPZ to remain silent (Ritter et al., 2018; Sunness et al., 2004) and found no evidence for large scale remapping of the visual cortex following late blindness (Baseler et al., 2011b; Goesaert et al., 2014; Haak et al., 2015). Masuda et al. (2008; 2010) found that the possibility to detect BOLD signals in LPZ depends on task and might reflect the upregulation or unmasking of feed-back projections from extrastriate areas to V1 that are normally suppressed in sighted individuals. These authors proposed that the presence of activations in LPZ might not need to reflect cortical reorganization. Similar conclusions were reached by studies that simulated scotomas in sighted humans artificially by removing the visual stimulus in a given location of a rich image. This studies found consistently altered receptive fields properties around the simulated scotoma (Binda et al., 2013; Dumoulin and Knapen, 2018; Haak et al., 2012). In sum, studies in patients with retinal dystrophies have reported mixed results regarding the presence of remapped visual activation in LPZ. Identifying the origin of such activations remains matter of open debate: it can reflect cortical reorganization leading to the formation of new connections or changes in the connections strength between neurons or to unmasking of existing connections normally suppressed in sighted individuals. 3.2 Cross-modal plasticity in blind individuals Another widely studied phenomena, considered a marker of experience-dependent neuroplasticity, is the cross-modal reorganization of the visual cortex in blind individuals: several studies observed occipital activations elicited by non-visual sensory stimulation in congenital and early-onset blinds (Amedi et al., 2010; Burton, 2003; Collignon et al., 2009, 2011b; Ptito et al., 2005). The mechanisms leading such cross-modal reorganization are still under debate. Some authors proposed that cross- modal activations in blinds reflect unmasking of 'latent' cross-modal connections that are normally suppressed (Merabet et al., 2007, 2008; Pascual-Leone and Hamilton, 2001), while others pointed at an additive shifts of the cross-modal responses in early- blind individuals rather than at rescaling or unmasking processes (Fine and Park, 2018; Lewis et al., 2010). Interestingly the cross-modal reorganization of typically visual areas is not random, but seems to rather reflect a supramodal functional organization of the brain, encoding for an abstract representation of the perceived stimuli independently from the sensory input (Pietrini et al., 2004; Ricciardi et al., 2014a,b; Ricciardi and Pietrini, 2011). For example, responses in the typically visual motion area MT+ can be elicited by motion-specific auditory and tactile stimulation in early-blinds and sight recovered individuals (Ricciardi et al., 2007; Saenz et al., 2008). However the overlap between cross-modal responses should be taken carefully and further investigated to exclude artefactual co-localizations of cross-modal responses within the same area (Jiang et al., 2015). Yet, signals reflecting functional selective cross- modal plasticity have been reported by several studies across different cortical areas and cognitive functions (such as object recognition, stimuli localization in the space, reading) and seems to be a general reorganizational principle of the brain independently from the deprived modality (Heimler et al., 2014): areas typically involved in spoken language processing for example, are recruited by sign language in deaf people (MacSweeney et al., 2008). The presence of cross-modal responses in late-onsets blind individuals is more controversial. For instance, Bedny et al. (2010) found activations elicited by auditory motion stimuli in hMT+ only in congenital, but not in late-blind subjects. Similarly the perceived direction of auditory moving stimuli could be classified in early, but not in late-blind individuals (Jiang et al., 2014, 2016) using multi voxel pattern analysis. However, there is also evidence for the cross-modal reorganization to take place when blindness develops late in life: although reduced with respect to those observed in early-blind patients, both auditory and tactile activations of the visual cortices have been described in late-blinds (Burton, 2003; Cunningham et al., 2011, 2015a,b; Voss et al., 2006), and one study showed that the extent and strength of tactile-evoked responses in V1 correlate with vision loss in late-blind individuals affected with retinitis pigmentosa (RP) (Cunningham et al., 2011, 2015b). The underlying anatomical pathways supporting cross-modal responses may differ between early and late-blind individuals (Collignon et al., 2009). For example, Collingon et al. (2013) found that the auditory activity in the occipital cortex of congenitally blind individuals was most likely conveyed by direct connections from A1 to V1, whereas in late -- blind individuals it appeared to be conveyed by feedback projections from multisensory parietal areas. The specific mechanisms underlying the rerouting of non-visual information are still unclear (Voss, 2019), however they might deeply differ depending on the age of blindness onset: while non-visual connections to the occipital cortex in congenital/early-blind individuals might be due to the lack of pruning typically triggered by visual experience, cross-modal responses in late-blind individuals might reflect the unmasking of non-visual pre-existing connections (normally supporting multisensory integration in sighted individuals) which can be progressively reinforced and result in permanent structural changes with new synapsis formation (Merabet et al., 2008). Whatever the cause or the precise mechanisms underlying cross-modal plasticity might be, it can potentially interfere with the outcome of restoration techniques (Collignon et al., 2011a; Merabet et al., 2005). This possibility has been clearly demonstrated in deaf children who underwent cochlear implant: sustained and prolonged periods of deafness induced stronger cross-modal reorganization of acoustic cortex (as measured by glucose hypometabolism) and hampered recovery after cochlear implant (Lee et al., 2001). Although the predictive link between the extent of cross-modal reorganization and successful outcome of vision restoration still has to be directly demonstrated, one potential prognostic predictor of treatment effectiveness might be V1 cortical thickness: Aguirre et al (2016) showed that this parameter is strictly related to the strengths of cross-modal responses, independently of the patient age and blinding pathology. Thus, cross-modal reorganization might pose a major challenge to vision restoration therapies, and if we take into account the lower plasticity in adulthood, it seem even more difficult to imagine successful visual recovery in late-blinds. On the other hand however, it is possible that sight restoration might be more feasible in late than in early- blind individuals, because visual cortex wiring successfully occurred before the disease in late -- blind individuals. A recent study showed that functional connectivity between visual areas is still retinotopically organized in blind patients affected with juvenile macular degeneration, even after prolonged period of visual deprivation, directing further hope on the possibility to have positive outcomes from sight restoration techniques that can rely on a relatively intact visual system (Haak et al., 2016). Yet, while these findings are encouraging, they do not guarantee that the adult visual brain retains the plastic potential necessary to mask the non-visual inputs potentiated or newly created during the blindness period and to boost the responses to the restored visual input; nor is it obvious that the visual system would 'learn to see again' with an artificial and monocular input. 3.3 Cortical plasticity after sight restoration The first attempts to restore vision in adulthood focused on early-blind patients and showed limited visual recovery in these patients (Dormal et al., 2015; Fine et al., 2003; Gregory and Wallace, 1963; Saenz et al., 2008). Patient MM became blind at the age of three years old, and once his sight was restored in his 40ies, he was only able to recover visual abilities strictly related to the visual experience before blindness, such as perception of simple forms, color and motion, while perception of more complex 3D forms, objects and faces remained severely impaired even after several years of restored vision (Fine et al., 2003; Huber et al., 2015). fMRI studies on patient MM and on another early-blind patient whose vision was partially restored in adulthood, showed that cortical plasticity was also limited although not completely absent: several months after vision restoration, cross-modal auditory responses continued to coexist with the restored visual activations in area V1 (Dormal et al., 2015) and MT (Saenz et al., 2008). Interestingly however Dormal et al. (2015) observed that cross-modal responses in extra striate areas decreased after surgery and vision improved suggesting that cross-modal reorganization can be partially reversed in their early-blind patient. Very few studies tracked the cortical reorganization process in late-blind patients after vision restoration. One study tracked the cortical responses in a 80 years old woman with wet age-related AMD undergoing intravitreal antiangiogenic injections (ranibizumab) over about 1 year period (Baseler et al., 2011a). Microperimetry showed that the scotoma decreased over time, visual acuity, fixation stability and reading skills improved as well. Interestingly after the first treatment, BOLD responses elicited by full-field flickering lights showed a tendency to be located in more posterior occipital regions, corresponding to the patient scotoma, however without filling it completely. Cunningham et al (2015a) tested two late-blind patients affected by retinitis pigmentosa (RP) implanted with Argus II Retinal Prosthesis and showed that the strength of tactile-evoked responses in V1 depended from the time from surgery: cross- modal activations were much reduced in the patient implanted for fifteen weeks before the scanning with respect to the patient who had the implant only for six weeks (Fig 2A). No visual responses were detected in this study, potentially due to the relatively short time period between the implantation and the scanning. A more recent study tested RP patients implanted with the same system at least six months after surgery and found that visual responses to flashes of lights increased in LGN and V1 after the surgery (Fig. 2B) (Castaldi et al., 2016). Importantly visual recovery (quantified as the behavioral performance on a challenging detection task) depended on the time from surgery and practice with the device (Fig. 2C) and was mirrored by enhanced BOLD responses to flashes after surgery, suggesting that the activity measured in visual areas had a functional relevance. Interestingly some weak visual activations were present already before the surgery, although participants never reported perceiving the flashing lights during the scanning. In particular, responses in extra striate areas were stronger before the surgery, whereas V1 showed stronger activations after the surgery. It is possible that even before surgery some spared visual input reached patients' V1, given the consistent albeit small BOLD response observed. However these visual signals might have been actively suppressed by extra striate areas, probably because the visual stimulation was not sufficiently reliable, aberrant and delayed. The suppression could have cross-sensory or motor origin, but also visual given that direct thalamic visual input to associative cortex have repeatedly observed in normal brain (Ajina et al., 2015; Tamietto and Morrone, 2016). Animal models reported spontaneous ganglion cells hyperactivity during photoreceptors degeneration process (Ivanova et al., 2016; Trenholm and Awatramani, 2015; Zeck, 2016). Suppressing such paroxysmal discharges might be beneficial faster cross-modal reorganization. This possibility is in line with the idea of a dysfunctional gating mechanism in blinds which would allow feedback projections to freely interact with the incoming input signal (Masuda et al., 2010), even when this is noisy and aberrant and the nature of this interaction might be inhibitory. to prompt Taken together the results of these experiments suggest that when visual signals are restored in late-blinds, not only cross-modal responses in V1 needs to be attenuated or eliminated, but also the suppression of the incoming visual input to V1 should be released, and both these processes might take a long time. Yet, t he decreased tactile-evoked responses (Cunningham et al., 2015a) and the increased visual BOLD signal in V1 (Castaldi et al., 2016) in patients that used the prosthesis for longer time, suggest that a spared plastic potential is retained by the adult visual brain and encourage continuative research to overcome the major obstacles limiting the expected outcome from vision restoration techniques. These obstacles include, among others, the limited quality of visual percepts that can be obtained (at best) with the current technology - for a review on advantages and limitations of different vision restoration methodologies see: Fine and Boynton (2015). For example, simulation of the likely perceived images with epiretinal devices showed that they might look extremely distorted because the electrode array stimulate unspecifically axons of the ganglion cells together with their cellular bodies (Fine and Boynton, 2015). It is thus not surprising that patients find particularly difficult to recover complex aspects of visual perception such as correctly discriminating the direction of drifting gratings (Castaldi et al., 2016). Although the technology can certainly be further improved, several studies have already shown that, after extensive training, RP patients implanted with retinal prosthesis can learn to perform some easy task, which can be nevertheless important for the patients' quality of life, such as moving independently in the space, locating large sources of light and even read large highly contrasted letters (Barry et al., 2012; Castaldi et al., 2016; Chader et al., 2009; da Cruz et al., 2013; Dorn et al., 2013; Rizzo et al., 2014). Improvement of visual acuity and visual field perimetry were reported in patients implanted with both epiretinal and subretinal implants (Chow 2013; Chow et al., 2004, 2010; Rizzo et al., 2014, 2015). Interestingly visual field improvement extended outside the retinotopic regions directly stimulated by the implant (Rizzo et al., 2014) and even in the unoperated eye (Rizzo et al., 2015). The fact that the visual field recovering is not strictly limited to the stimulation site might suggest that the neuroplastic response acts 'peripherally' and originate from the release of retinal trophic factors induced by the current injection which diffuse to non-stimulated regions of the retina (Ciavatta et al., 2009). However it is also possible that the artificial vision provided by retinal implants reopens visual plasticity at 'central level', which would better explain the visual recovery of the fellow eye observed at least in one study (Rizzo et al., 2015). Perhaps the fellow eye benefits from an anterograde effect mediated by cross talk of neural discharges along the optic nerve after the chiasma or at the level of LGN where the projections from the two eyes are closely interlayered. Interestingly some increase of BOLD response have also been observed at LGN level after prolonged use of the prosthetic devise (Castaldi et al., 2016). In sum, despite long time and extensive training is needed to recover a functional use of artificial vision and although the specific source of such neuroplastic responses is currently unknown, the reviewed results suggest that vision restoration techniques can rely on residual neuroplasticity retained by the adult brain and that, especially for late-blinds, it might be possible to restore vision even after several years of blindness. 4. Conclusions We reviewed behavioral and neural evidence suggesting that a considerable degree of neuroplasticity is preserved well beyond the closure of the critical periods for vision. Behavioral and neural evidence indicates that in sighted individual short periods of monocular deprivation can trigger homeostatic plasticity and that this strategy can be successfully used to improve visual perception in amblyopic adults. Evidence for cross- modal reorganization of the visual brain of blind individuals is compelling and recent findings suggest that this process can be reverted even after several years of blindness. However it is important to acknowledge the fact that this neuroplasticity is a very slow process and that the quality of the regained visual perception is limited. Yet, the field of visual restoration techniques is still at the beginning and much improvement can be expected in the upcoming years. Certainly neuroscientists can contribute to the development of this field by exploring the properties, characteristics and mechanisms of the spared plastic potential retained by the human visual brain to calibrate optimally vision restoration therapies and maximize their expected outcome. Perhaps, the most outstanding question to answer is whether we can boost neuroplasticity to make visual recovery more effective and rapid. Although the answer might depend on the age of blindness onset and on the cause determining the visual deficit, identifying some common principles and mechanisms guiding neuroplasticity during development and adulthood and across modalities mighty lead to develop strategies effective under a wide range of circumstances and diseases. Hopefully in the next decades we will have the complete approach to rescue visual function, facilitating the adult brain to learn to "see" again, even after several years of blindness or abnormal visual experience. Acknowledgments This research has received funding from Fondazione Roma under the Grants for Biomedical Research: Retinitis Pigmentosa (RP)-Call for proposals 2013 (http:// www.fondazioneroma.it/it/index.html, http://wf-fondazioneroma.cbim.it/), project title: Cortical Plasticity in Retinitis Pigmentosa: an Integrated Study from Animal Models to Humans, from the European Research Council under the European Union's Seventh Framework Programme (FPT/2007-2013) under grant agreement No. 338866 ECPLAIN (http://www.pisavisionlab.org/index.php/ projects/ecsplain) and under the European Union's Horizon 2020 research and innovation programme (grant agreement No 801715 -- PUPILTRAITS). References Abbott, L.F., Nelson, S.B., 2000. Synaptic plasticity: taming the beast. Nature neuroscience 3 Suppl, 1178-1183. Aguirre, G.K., Datta, R., Benson, N.C., Prasad, S., Jacobson, S.G., Cideciyan, A.V., Bridge, H., Watkins, K.E., Butt, O.H., Dain, A.S., Brandes, L., Gennatas, E.D., 2016. Patterns of Individual Variation in Visual Pathway Structure and Function in the Sighted and Blind. Plos One 11, e0164677. Ajina, S., Pestilli, F., Rokem, A., Kennard, C., Bridge, H., 2015. Human blindsight is mediated by an intact geniculo-extrastriate pathway. eLife 4. Amedi, A., Raz, N., Azulay, H., Malach, R., Zohary, E., 2010. Cortical activity during tactile exploration of objects in blind and sighted humans. Restor Neurol Neurosci 28, 143-156. Bai, J., Dong, X., He, S., Bao, M., 2017. Monocular deprivation of Fourier phase information boosts the deprived eye's dominance during interocular competition but not interocular phase combination. Neuroscience 352, 122-130. Baker, C.I., Dilks, D.D., Peli, E., Kanwisher, N., 2008. Reorganization of visual processing in macular degeneration: replication and clues about the role of foveal loss. Vision Res 48, 1910-1919. Baker, C.I., Peli, E., Knouf, N., Kanwisher, N.G., 2005. Reorganization of visual processing in macular degeneration. The Journal of neuroscience : the official journal of the Society for Neuroscience 25, 614-618. Baroncelli, L., Bonaccorsi, J., Milanese, M., Bonifacino, T., Giribaldi, F., Manno, I., Cenni, M.C., Berardi, N., Bonanno, G., Maffei, L., Sale, A., 2012. Enriched experience and recovery from amblyopia in adult rats: impact of motor, social and sensory components. Neuropharmacology 62, 2388-2397. Baroncelli, L., Sale, A., Viegi, A., Maya Vetencourt, J.F., De Pasquale, R., Baldini, S., Maffei, L., 2010. Experience-dependent reactivation of ocular dominance plasticity in the adult visual cortex. Experimental neurology 226, 100-109. Barry, M.P., Dagnelie, G., Argus, I.I.S.G., 2012. Use of the Argus II retinal prosthesis to improve visual guidance of fine hand movements. Invest Ophthalmol Vis Sci 53, 5095-5101. Baseler, H.A., Gouws, A., Crossland, M.D., Leung, C., Tufail, A., Rubin, G.S., Morland, A.B., 2011a. Objective visual assessment of antiangiogenic treatment for wet age-related macular degeneration. Optometry and vision science : official publication of the American Academy of Optometry 88, 1255- 1261. Baseler, H.A., Gouws, A., Haak, K.V., Racey, C., Crossland, M.D., Tufail, A., Rubin, G.S., Cornelissen, F.W., Morland, A.B., 2011b. Large-scale remapping of visual cortex is absent in adult humans with macular degeneration. Nature neuroscience 14, 649-655. Bedny, M., Konkle, T., Pelphrey, K., Saxe, R., Pascual-Leone, A., 2010. Sensitive period for a multimodal response in human visual motion area MT/MST. Curr Biol 20, 1900-1906. Berardi, N., Pizzorusso, T., Maffei, L., 2000. Critical periods during sensory development. Curr Opin Neurobiol 10, 138-145. Berardi, N., Pizzorusso, T., Ratto, G.M., Maffei, L., 2003. Molecular basis of plasticity in the visual cortex. Trends in neurosciences 26, 369-378. Berardi, N., Sale, A., Maffei, L., 2015. Brain structural and functional development: genetics and experience. Developmental medicine and child neurology 57 Suppl 2, 4-9. Beyeler, M., Rokem, A., Boynton, G.M., Fine, I., 2017. Learning to see again: biological constraints on cortical plasticity and the implications for sight restoration technologies. Journal of neural engineering 14, 051003. Binda, P., Kurzawski, J.W., Lunghi, C., Biagi, L., Tosetti, M., Morrone, M.C., 2018. Response to short- term deprivation of the human adult visual cortex measured with 7T BOLD. eLife 7. Binda, P., Lunghi, C., 2017. Short-Term Monocular Deprivation Enhances Physiological Pupillary Oscillations. Neural plasticity 2017, 6724631. Binda, P., Thomas, J.M., Boynton, G.M., Fine, I., 2013. Minimizing biases in estimating the reorganization of human visual areas with BOLD retinotopic mapping. Journal of vision 13, 13. Blake, R., Logothetis, N., 2002. Visual competition. Nat Rev Neurosci 3, 13-21. Boroojerdi, B., Battaglia, F., Muellbacher, W., Cohen, L.G., 2001. Mechanisms underlying rapid experience-dependent plasticity in the human visual cortex. Proceedings of the National Academy of Sciences of the United States of America 98, 14698-14701. Boroojerdi, B., Bushara, K.O., Corwell, B., Immisch, I., Battaglia, F., Muellbacher, W., Cohen, L.G., 2000. Enhanced excitability of the human visual cortex induced by short-term light deprivation. Cerebral cortex 10, 529-534. Braddick, O., Atkinson, J., 2011. Development of human visual function. Vision Res 51, 1588-1609. Burton, H., 2003. Visual cortex activity in early and late blind people. The Journal of neuroscience : the official journal of the Society for Neuroscience 23, 4005-4011. Calford, M.B., Chino, Y.M., Das, A., Eysel, U.T., Gilbert, C.D., Heinen, S.J., Kaas, J.H., Ullman, S., 2005. Neuroscience: rewiring the adult brain. Nature 438, E3; discussion E3-4. Carandini, M., Heeger, D.J., 2011. Normalization as a canonical neural computation. Nat Rev Neurosci 13, 51-62. Castaldi, E., Cicchini, G.M., Cinelli, L., Biagi, L., Rizzo, S., Morrone, M.C., 2016. Visual BOLD Response in Late Blind Subjects with Argus II Retinal Prosthesis. PLoS Biol 14, e1002569. Chader, G.J., Weiland, J., Humayun, M.S., 2009. Artificial vision: needs, functioning, and testing of a retinal electronic prosthesis. Progress in brain research 175, 317-332. Chow, A.Y., 2013. Retinal Prostheses Development in Retinitis Pigmentosa Patients-Progress and Comparison. Asia-Pacific journal of ophthalmology 2, 253-268. Chow, A.Y., Bittner, A.K., Pardue, M.T., 2010. The artificial silicon retina in retinitis pigmentosa patients (an American Ophthalmological Association thesis). Transactions of the American Ophthalmological Society 108, 120-154. Chow, A.Y., Chow, V.Y., Packo, K.H., Pollack, J.S., Peyman, G.A., Schuchard, R., 2004. The artificial silicon retina microchip for the treatment of vision loss from retinitis pigmentosa. Archives of ophthalmology 122, 460-469. Ciavatta, V.T., Kim, M., Wong, P., Nickerson, J.M., Shuler, R.K., Jr., McLean, G.Y., Pardue, M.T., 2009. Retinal expression of Fgf2 in RCS rats with subretinal microphotodiode array. Invest Ophthalmol Vis Sci 50, 4523-4530. Collignon, O., Champoux, F., Voss, P., Lepore, F., 2011a. Sensory rehabilitation in the plastic brain. Progress in brain research 191, 211-231. Collignon, O., Dormal, G., Albouy, G., Vandewalle, G., Voss, P., Phillips, C., Lepore, F., 2013. Impact of blindness onset on the functional organization and the connectivity of the occipital cortex. Brain : a journal of neurology 136, 2769-2783. Collignon, O., Vandewalle, G., Voss, P., Albouy, G., Charbonneau, G., Lassonde, M., Lepore, F., 2011b. Functional specialization for auditory-spatial processing in the occipital cortex of congenitally blind humans. Proceedings of the National Academy of Sciences of the United States of America 108, 4435- 4440. Collignon, O., Voss, P., Lassonde, M., Lepore, F., 2009. Cross-modal plasticity for the spatial processing of sounds in visually deprived subjects. Experimental brain research 192, 343-358. Cunningham, S.I., Shi, Y., Weiland, J.D., Falabella, P., Olmos de Koo, L.C., Zacks, D.N., Tjan, B.S., 2015a. Feasibility of Structural and Functional MRI Acquisition with Unpowered Implants in Argus II Retinal Prosthesis Patients: A Case Study. Transl Vis Sci Technol 4, 6. Cunningham, S.I., Weiland, J.D., Bao, P., Lopez-Jaime, G.R., Tjan, B.S., 2015b. Correlation of vision loss with tactile-evoked V1 responses in retinitis pigmentosa. Vision Res 111, 197-207. Cunningham, S.I., Weiland, J.D., Bao, P., Tjan, B.S., 2011. Visual cortex activation induced by tactile stimulation in late-blind individuals with retinitis pigmentosa. Conference proceedings : ... Annual International Conference of the IEEE Engineering in Medicine and Biology Society. IEEE Engineering in Medicine and Biology Society. Annual Conference 2011, 2841-2844. da Cruz, L., Coley, B.F., Dorn, J., Merlini, F., Filley, E., Christopher, P., Chen, F.K., Wuyyuru, V., Sahel, J., Stanga, P., Humayun, M., Greenberg, R.J., Dagnelie, G., Argus, I.I.S.G., 2013. The Argus II epiretinal prosthesis system allows letter and word reading and long-term function in patients with profound vision loss. The British journal of ophthalmology 97, 632-636. Di Russo, F., Martinez, A., Sereno, M.I., Pitzalis, S., Hillyard, S.A., 2002. Cortical sources of the early components of the visual evoked potential. Human brain mapping 15, 95-111. Diamond, J.P., 2001. The Pupil. Anatomy, Physiology and Clinical Applications. The British journal of ophthalmology 85, 121E. Dilks, D.D., Baker, C.I., Peli, E., Kanwisher, N., 2009. Reorganization of visual processing in macular degeneration is not specific to the "preferred retinal locus". The Journal of neuroscience : the official journal of the Society for Neuroscience 29, 2768-2773. Dilks, D.D., Julian, J.B., Peli, E., Kanwisher, N., 2014. Reorganization of visual processing in age-related macular degeneration depends on foveal loss. Optometry and vision science : official publication of the American Academy of Optometry 91, e199-206. Donahue, M.J., Near, J., Blicher, J.U., Jezzard, P., 2010. Baseline GABA concentration and fMRI response. NeuroImage 53, 392-398. Dormal, G., Lepore, F., Harissi-Dagher, M., Albouy, G., Bertone, A., Rossion, B., Collignon, O., 2015. Tracking the evolution of crossmodal plasticity and visual functions before and after sight restoration. J Neurophysiol 113, 1727-1742. Dorn, J.D., Ahuja, A.K., Caspi, A., da Cruz, L., Dagnelie, G., Sahel, J.A., Greenberg, R.J., McMahon, M.J., gus, I.I.S.G., 2013. The Detection of Motion by Blind Subjects With the Epiretinal 60-Electrode (Argus II) Retinal Prosthesis. JAMA ophthalmology 131, 183-189. Dosher, B., Lu, Z.L., 2017. Visual Perceptual Learning and Models. Annu Rev Vis Sci 3, 343-363. Dumoulin, S.O., Knapen, T., 2018. How Visual Cortical Organization Is Altered by Ophthalmologic and Neurologic Disorders. Annu Rev Vis Sci 4, 357-379. Fagiolini, M., Fritschy, J.M., Low, K., Mohler, H., Rudolph, U., Hensch, T.K., 2004. Specific GABAA circuits for visual cortical plasticity. Science 303, 1681-1683. Fagiolini, M., Hensch, T.K., 2000. Inhibitory threshold for critical-period activation in primary visual cortex. Nature 404, 183-186. Ferreira, S., Pereira, A.C., Quendera, B., Reis, A., Silva, E.D., Castelo-Branco, M., 2017. Primary visual cortical remapping in patients with inherited peripheral retinal degeneration. Neuroimage Clin 13, 428- 438. Fierro, B., Brighina, F., Vitello, G., Piazza, A., Scalia, S., Giglia, G., Daniele, O., Pascual-Leone, A., 2005. Modulatory effects of low- and high-frequency repetitive transcranial magnetic stimulation on visual cortex of healthy subjects undergoing light deprivation. The Journal of physiology 565, 659-665. Fine, I., Boynton, G.M., 2015. Pulse trains to percepts: the challenge of creating a perceptually intelligible world with sight recovery technologies. Philosophical transactions of the Royal Society of London. Series B, Biological sciences 370, 20140208. Fine, I., Park, J.M., 2018. Blindness and Human Brain Plasticity. Annu Rev Vis Sci 4, 337-356. Fine, I., Smallman, H.S., Doyle, P., MacLeod, D.I., 2002. Visual function before and after the removal of bilateral congenital cataracts in adulthood. Vision Res 42, 191-210. Fine, I., Wade, A.R., Brewer, A.A., May, M.G., Goodman, D.F., Boynton, G.M., Wandell, B.A., MacLeod, D.I., 2003. Long-term deprivation affects visual perception and cortex. Nature neuroscience 6, 915- 916. Fiorentini, A., Berardi, N., 1980. Perceptual learning specific for orientation and spatial frequency. Nature 287, 43-44. Fronius, M., Cirina, L., Ackermann, H., Kohnen, T., Diehl, C.M., 2014. Efficiency of electronically monitored amblyopia treatment between 5 and 16 years of age: new insight into declining susceptibility of the visual system. Vision Res 103, 11-19. Fuchs, E., Flugge, G., 2014. Adult neuroplasticity: more than 40 years of research. Neural plasticity 2014, 541870. Gilbert, C.D., Wiesel, T.N., 1992. Receptive field dynamics in adult primary visual cortex. Nature 356, 150-152. Goesaert, E., Van Baelen, M., Spileers, W., Wagemans, J., Op de Beeck, H.P., 2014. Visual Space and Object Space in the Cerebral Cortex of Retinal Disease Patients. Plos One 9. Gregory, R., Wallace, J., 1963. Recovery from early blindness: a case study. Heffer, Cambridge. Haak, K.V., Cornelissen, F.W., Morland, A.B., 2012. Population receptive field dynamics in human visual cortex. Plos One 7, e37686. Haak, K.V., Morland, A.B., Engel, S.A., 2015. Plasticity, and Its Limits, in Adult Human Primary Visual Cortex. Multisensory research 28, 297-307. Haak, K.V., Morland, A.B., Rubin, G.S., Cornelissen, F.W., 2016. Preserved retinotopic brain connectivity in macular degeneration. Ophthalmic & physiological optics : the journal of the British College of Ophthalmic Opticians 36, 335-343. Hanover, J.L., Huang, Z.J., Tonegawa, S., Stryker, M.P., 1999. Brain-derived neurotrophic factor overexpression induces precocious critical period in mouse visual cortex. The Journal of neuroscience : the official journal of the Society for Neuroscience 19, RC40. Harauzov, A., Spolidoro, M., DiCristo, G., De Pasquale, R., Cancedda, L., Pizzorusso, T., Viegi, A., Berardi, N., Maffei, L., 2010. Reducing intracortical inhibition in the adult visual cortex promotes ocular dominance plasticity. The Journal of neuroscience : the official journal of the Society for Neuroscience 30, 361-371. He, H.Y., Hodos, W., Quinlan, E.M., 2006. Visual deprivation reactivates rapid ocular dominance plasticity in adult visual cortex. The Journal of neuroscience : the official journal of the Society for Neuroscience 26, 2951-2955. Heeger, D.J., 1992. Normalization of cell responses in cat striate cortex. Visual neuroscience 9, 181- 197. Heimel, J.A., van Versendaal, D., Levelt, C.N., 2011. The role of GABAergic inhibition in ocular dominance plasticity. Neural plasticity 2011, 391763. Heimler, B., Weisz, N., Collignon, O., 2014. Revisiting the adaptive and maladaptive effects of crossmodal plasticity. Neuroscience 283, 44-63. Hensch, T.K., Fagiolini, M., 2005. Excitatory-inhibitory balance and critical period plasticity in developing visual cortex. Progress in brain research 147, 115-124. Hensch, T.K., Fagiolini, M., Mataga, N., Stryker, M.P., Baekkeskov, S., Kash, S.F., 1998. Local GABA circuit control of experience-dependent plasticity in developing visual cortex. Science 282, 1504-1508. Hensch, T.K., Quinlan, E.M., 2018. Critical periods in amblyopia. Visual neuroscience 35, E014. Huang, Z.J., Kirkwood, A., Pizzorusso, T., Porciatti, V., Morales, B., Bear, M.F., Maffei, L., Tonegawa, S., 1999. BDNF regulates the maturation of inhibition and the critical period of plasticity in mouse visual cortex. Cell 98, 739-755. Hubel, D.H., Wiesel, T.N., 1970. The period of susceptibility to the physiological effects of unilateral eye closure in kittens. The Journal of physiology 206, 419-436. Hubel, D.H., Wiesel, T.N., LeVay, S., 1977. Plasticity of ocular dominance columns in monkey striate cortex. Philosophical transactions of the Royal Society of London. Series B, Biological sciences 278, 377- 409. Huber, E., Webster, J.M., Brewer, A.A., MacLeod, D.I., Wandell, B.A., Boynton, G.M., Wade, A.R., Fine, I., 2015. A lack of experience-dependent plasticity after more than a decade of recovered sight. Psychological science 26, 393-401. Ivanova, E., Yee, C.W., Baldoni, R., Sagdullaev, B.T., 2016. Aberrant activity in retinal degeneration impairs central visual processing and relies on Cx36-containing gap junctions. Exp Eye Res 150, 81-89. Jiang, F., Beauchamp, M.S., Fine, I., 2015. Re-examining overlap between tactile and visual motion responses within hMT+ and STS. NeuroImage 119, 187-196. Jiang, F., Stecker, G.C., Boynton, G.M., Fine, I., 2016. Early Blindness Results in Developmental Plasticity for Auditory Motion Processing within Auditory and Occipital Cortex. Frontiers in human neuroscience 10, 324. Jiang, F., Stecker, G.C., Fine, I., 2014. Auditory motion processing after early blindness. Journal of vision 14, 4. Joshi, S., Li, Y., Kalwani, R.M., Gold, J.I., 2016. Relationships between Pupil Diameter and Neuronal Activity in the Locus Coeruleus, Colliculi, and Cingulate Cortex. Neuron 89, 221-234. Kasamatsu, T., Pettigrew, J.D., Ary, M., 1979. Restoration of visual cortical plasticity by local microperfusion of norepinephrine. The Journal of comparative neurology 185, 163-181. Kasamatsu, T., Pettigrew, J.D., Ary, M., 1981. Cortical recovery from effects of monocular deprivation: acceleration with norepinephrine and suppression with 6-hydroxydopamine. J Neurophysiol 45, 254- 266. Kiorpes, L., Kiper, D.C., O'Keefe, L.P., Cavanaugh, J.R., Movshon, J.A., 1998. Neuronal correlates of amblyopia in the visual cortex of macaque monkeys with experimental strabismus and anisometropia. The Journal of neuroscience : the official journal of the Society for Neuroscience 18, 6411-6424. Lee, D.S., Lee, J.S., Oh, S.H., Kim, S.K., Kim, J.W., Chung, J.K., Lee, M.C., Kim, C.S., 2001. Cross-modal plasticity and cochlear implants. Nature 409, 149-150. Levelt, C.N., Hubener, M., 2012. Critical-period plasticity in the visual cortex. Annual review of neuroscience 35, 309-330. Levelt, W.J., 1965. On Binocular Rivalry Soesterberg: Netherlands: Institution for Perception. Levi, D.M., Carkeet, A., 1993. Amblyopia: a consequence of abnormal visual development. , in: Simons, K. (Ed.), Early Visual Development, Normal and Abnormal. Oxford University Press, New York, NY, pp. 391 -- 408. Lewis, L.B., Saenz, M., Fine, I., 2010. Mechanisms of cross-modal plasticity in early-blind subjects. J Neurophysiol 104, 2995-3008. Lou, A.R., Madsen, K.H., Paulson, O.B., Julian, H.O., Prause, J.U., Siebner, H.R., Kjaer, T.W., 2011. Monocular visual deprivation suppresses excitability in adult human visual cortex. Cerebral cortex 21, 2876-2882. Lunghi, C., Berchicci, M., Morrone, M.C., Di Russo, F., 2015a. Short-term monocular deprivation alters early components of visual evoked potentials. The Journal of physiology 593, 4361-4372. Lunghi, C., Burr, D.C., Morrone, C., 2011. Brief periods of monocular deprivation disrupt ocular balance in human adult visual cortex. Curr Biol 21, R538-539. Lunghi, C., Burr, D.C., Morrone, M.C., 2013. Long-term effects of monocular deprivation revealed with binocular rivalry gratings modulated in luminance and in color. Journal of vision 13. Lunghi, C., Emir, U.E., Morrone, M.C., Bridge, H., 2015b. Short-term monocular deprivation alters GABA in the adult human visual cortex. Curr Biol 25, 1496-1501. Lunghi, C., Sale, A., 2015. A cycling lane for brain rewiring. Curr Biol 25, R1122-1123. Lunghi, C., Sframeli, T., Lepri, A., Lepri, M., Lisi, D., Sale, A., Morrone, M.C., 2018. A new counterintuitive training for adult amblyopia. Annals of Clinical and Translational Neurology. MacSweeney, M., Capek, C.M., Campbell, R., Woll, B., 2008. The signing brain: the neurobiology of sign language. Trends in cognitive sciences 12, 432-440. Maffei, A., Turrigiano, G., 2008a. The age of plasticity: developmental regulation of synaptic plasticity in neocortical microcircuits. Progress in brain research 169, 211-223. Maffei, A., Turrigiano, G.G., 2008b. Multiple modes of network homeostasis in visual cortical layer 2/3. The Journal of neuroscience : the official journal of the Society for Neuroscience 28, 4377-4384. Malenka, R.C., Bear, M.F., 2004. LTP and LTD: an embarrassment of riches. Neuron 44, 5-21. Masuda, Y., Dumoulin, S.O., Nakadomari, S., Wandell, B.A., 2008. V1 projection zone signals in human macular degeneration depend on task, not stimulus. Cerebral cortex 18, 2483-2493. Masuda, Y., Horiguchi, H., Dumoulin, S.O., Furuta, A., Miyauchi, S., Nakadomari, S., Wandell, B.A., 2010. Task-dependent V1 responses in human retinitis pigmentosa. Invest Ophthalmol Vis Sci 51, 5356-5364. Maurer, D., Lewis, T.L., Mondloch, C.J., 2005. Missing sights: consequences for visual cognitive development. Trends in cognitive sciences 9, 144-151. Maurer, D., Mondloch, C.J., Lewis, T.L., 2007. Effects of early visual deprivation on perceptual and cognitive development. Progress in brain research 164, 87-104. Maya Vetencourt, J.F., Sale, A., Viegi, A., Baroncelli, L., De Pasquale, R., O'Leary, O.F., Castren, E., Maffei, L., 2008. The antidepressant fluoxetine restores plasticity in the adult visual cortex. Science 320, 385-388. Merabet, L.B., Hamilton, R., Schlaug, G., Swisher, J.D., Kiriakopoulos, E.T., Pitskel, N.B., Kauffman, T., Pascual-Leone, A., 2008. Rapid and reversible recruitment of early visual cortex for touch. Plos One 3, e3046. Merabet, L.B., Rizzo, J.F., Amedi, A., Somers, D.C., Pascual-Leone, A., 2005. What blindness can tell us about seeing again: merging neuroplasticity and neuroprostheses. Nat Rev Neurosci 6, 71-77. Merabet, L.B., Swisher, J.D., McMains, S.A., Halko, M.A., Amedi, A., Pascual-Leone, A., Somers, D.C., 2007. Combined activation and deactivation of visual cortex during tactile sensory processing. J Neurophysiol 97, 1633-1641. Morishita, H., Miwa, J.M., Heintz, N., Hensch, T.K., 2010. Lynx1, a cholinergic brake, limits plasticity in adult visual cortex. Science 330, 1238-1240. Muthukumaraswamy, S.D., Edden, R.A., Jones, D.K., Swettenham, J.B., Singh, K.D., 2009. Resting GABA concentration predicts peak gamma frequency and fMRI amplitude in response to visual stimulation in humans. Proceedings of the National Academy of Sciences of the United States of America 106, 8356-8361. Ooi, T.L., He, Z.J., 2001. Sensory eye dominance. Optometry 72, 168-178. Pascual-Leone, A., Amedi, A., Fregni, F., Merabet, L.B., 2005. The plastic human brain cortex. Annual review of neuroscience 28, 377-401. Pascual-Leone, A., Hamilton, R., 2001. The metamodal organization of the brain. Progress in brain research 134, 427-445. Pietrini, P., Furey, M.L., Ricciardi, E., Gobbini, M.I., Wu, W.H., Cohen, L., Guazzelli, M., Haxby, J.V., 2004. Beyond sensory images: Object-based representation in the human ventral pathway. Proceedings of the National Academy of Sciences of the United States of America 101, 5658-5663. Pitskel, N.B., Merabet, L.B., Ramos-Estebanez, C., Kauffman, T., Pascual-Leone, A., 2007. Time- dependent changes in cortical excitability after prolonged visual deprivation. Neuroreport 18, 1703- 1707. Pizzorusso, T., Medini, P., Berardi, N., Chierzi, S., Fawcett, J.W., Maffei, L., 2002. Reactivation of ocular dominance plasticity in the adult visual cortex. Science 298, 1248-1251. Ptito, M., Moesgaard, S.M., Gjedde, A., Kupers, R., 2005. Cross-modal plasticity revealed by electrotactile stimulation of the tongue in the congenitally blind. Brain : a journal of neurology 128, 606-614. Ramamurthy, M., Blaser, E., 2018. Assessing the kaleidoscope of monocular deprivation effects. Journal of vision 18, 14. Ricciardi, E., Bonino, D., Pellegrini, S., Pietrini, P., 2014a. Mind the blind brain to understand the sighted one! Is there a supramodal cortical functional architecture? Neuroscience and biobehavioral reviews 41, 64-77. Ricciardi, E., Pietrini, P., 2011. New light from the dark: what blindness can teach us about brain function. Curr Opin Neurol 24, 357-363. Ricciardi, E., Tozzi, L., Leo, A., Pietrini, P., 2014b. Modality dependent cross-modal functional reorganization following congenital visual deprivation within occipital areas: a meta-analysis of tactile and auditory studies. Multisensory research 27, 247-262. Ricciardi, E., Vanello, N., Sani, L., Gentili, C., Scilingo, E.P., Landini, L., Guazzelli, M., Bicchi, A., Haxby, J.V., Pietrini, P., 2007. The effect of visual experience on the development of functional architecture in hMT+. Cerebral cortex 17, 2933-2939. Ritter, M., Hummer, A., Ledolter, A., Holder, G., Windischberger, C., Schmidt-Erfurth, U., 2018. Correspondence between functional and morphological assessment of retinal disease. Br J Ophthalmol. Rizzo, S., Belting, C., Cinelli, L., Allegrini, L., 2015. Visual field changes following implantation of the Argus II retinal prosthesis. Graefe's archive for clinical and experimental ophthalmology = Albrecht von Graefes Archiv fur klinische und experimentelle Ophthalmologie 253, 323-325. retinotopic cortical mapping and conventional Rizzo, S., Belting, C., Cinelli, L., Allegrini, L., Genovesi-Ebert, F., Barca, F., di Bartolo, E., 2014. The Argus II Retinal Prosthesis: 12-month outcomes from a single-study center. American journal of ophthalmology 157, 1282-1290. Saenz, M., Lewis, L.B., Huth, A.G., Fine, I., Koch, C., 2008. Visual Motion Area MT+/V5 Responds to Auditory Motion in Human Sight-Recovery Subjects. The Journal of neuroscience : the official journal of the Society for Neuroscience 28, 5141-5148. Sale, A., Berardi, N., Maffei, L., 2014. Environment and brain plasticity: towards an endogenous pharmacotherapy. Physiological reviews 94, 189-234. Sale, A., Maya Vetencourt, J.F., Medini, P., Cenni, M.C., Baroncelli, L., De Pasquale, R., Maffei, L., 2007. Environmental enrichment in adulthood promotes amblyopia recovery through a reduction of intracortical inhibition. Nature neuroscience 10, 679-681. Schumacher, E.H., Jacko, J.A., Primo, S.A., Main, K.L., Moloney, K.P., Kinzel, E.N., Ginn, J., 2008. Reorganization of visual processing is related to eccentric viewing in patients with macular degeneration. Restor Neurol Neurosci 26, 391-402. Sereno, M.I., 2005. Neuroscience: plasticity and its limits. Nature 435, 288-289. Smirnakis, S.M., Brewer, A.A., Schmid, M.C., Tolias, A.S., Schuz, A., Augath, M., Inhoffen, W., Wandell, B.A., Logothetis, N.K., 2005. Lack of long-term cortical reorganization after macaque retinal lesions. Nature 435, 300-307. Smith, G.B., Heynen, A.J., Bear, M.F., 2009. Bidirectional synaptic mechanisms of ocular dominance plasticity in visual cortex. Philosophical transactions of the Royal Society of London. Series B, Biological sciences 364, 357-367. Speed, H.D., Morrone, M.C., Burr, D.C., 1991. Effects of monocular deprivation on the development of visual inhibitory interactions in kittens. Visual neuroscience 7, 335-343. Spolidoro, M., Sale, A., Berardi, N., Maffei, L., 2009. Plasticity in the adult brain: lessons from the visual system. Experimental brain research 192, 335-341. Stryker, M.P., 2014. A Neural Circuit That Controls Cortical State, Plasticity, and the Gain of Sensory Responses in Mouse. Cold Spring Harbor symposia on quantitative biology 79, 1-9. Sunness, J.S., Liu, T., Yantis, S., 2004. Retinotopic mapping of the visual cortex using functional magnetic resonance imaging in a patient with central scotomas from atrophic macular degeneration. Ophthalmology 111, 1595-1598. Tamietto, M., Morrone, M.C., 2016. Visual Plasticity: Blindsight Bridges Anatomy and Function in the Visual System. Curr Biol 26, R70-R73. Trenholm, S., Awatramani, G.B., 2015. Origins of spontaneous activity in the degenerating retina. Front Cell Neurosci 9, 277. Turrigiano, G., 2012. Homeostatic synaptic plasticity: local and global mechanisms for stabilizing neuronal function. Cold Spring Harbor perspectives in biology 4, a005736. Turrigiano, G., Nelson, S.B., 2000. Hebb and homeostasis in neuronal plasticity. Curr Opin Neurobiol 10, 358-364. Voss, P., 2019. Brain (re)organization following visual loss. Wiley interdisciplinary reviews. Cognitive science 10, e1468. Voss, P., Gougoux, F., Lassonde, M., Zatorre, R.J., Lepore, F., 2006. A positron emission tomography study during auditory localization by late-onset blind individuals. Neuroreport 17, 383-388. Wandell, B.A., Smirnakis, S.M., 2009. Plasticity and stability of visual field maps in adult primary visual cortex. Nat Rev Neurosci 10, 873-884. Watanabe, T., Sasaki, Y., 2015. Perceptual learning: toward a comprehensive theory. Annual review of psychology 66, 197-221. Webber, A.L., Wood, J., 2005. Amblyopia: prevalence, natural history, functional effects and treatment. Clin Exp Optom 88, 365-375. Wiesel, T.N., Hubel, D.H., 1963. Effects of Visual Deprivation on Morphology and Physiology of Cells in the Cats Lateral Geniculate Body. J Neurophysiol 26, 978-993. Wiesel, T.N., Hubel, D.H., 1965a. Comparison of the effects of unilateral and bilateral eye closure on cortical unit responses in kittens. J Neurophysiol 28, 1029-1040. Wiesel, T.N., Hubel, D.H., 1965b. Extent of recovery from the effects of visual deprivation in kittens. J Neurophysiol 28, 1060-1072. Zeck, G., 2016. Aberrant Activity in Degenerated Retinas Revealed by Electrical Imaging. Front Cell Neurosci 10, 25. Zhou, J., Clavagnier, S., Hess, R.F., 2013. Short-term monocular deprivation strengthens the patched eye's contribution to binocular combination. Journal of vision 13. Zhou, J., Reynaud, A., Hess, R.F., 2014. Real-time modulation of perceptual eye dominance in humans. Proceedings. Biological sciences 281. Figure 1. Short-term visual deprivation induces functional reorganization of cortical circuits in sighted adult humans. (A) During 7T fMRI scanning participants' eyes were separately stimulated with band- pass noise and BOLD responses were measured before and after 2h of monocular deprivation. The flat map shows extensive BOLD response before deprivation elicited by the noise stimuli in all visual areas. (B) The percentage BOLD signal measured in V1 elicited by stimulation of the deprived eye and non-deprived eye was comparable before deprivation, while respectively enhanced and reduced after 2h of monocular deprivation. (C) The neural changes described are reflected at behavioral level: mean phase duration during binocular rivalry are balanced across the two eyes before deprivation, while mean phases are the deprived/non-deprived eye respectively. (D) These perceptual effects correlate with the deprivation index measured with fMRI. (A-D) Reproduced from Figure 1 and Figure 3 (Binda et al., 2018), eLife, published under the Creative Commons Attribution 4.0 International 4.0; https://creativecommons.org/licenses/by/4.0/)." longer/shorter after deprivation (CC for Public License BY Figure 2. After sufficient time and practice the adult visual brain can learn to see again with artificial vision. (A) BOLD responses elicited by three different tactile tasks in two RP patients implanted with Argus II retinal prosthesis (subjects A1 and A2) who underwent fMRI scanning after 6 and 15 weeks after surgery. Participants were required to haptically determine the symmetry of raised-line shapes (shape task), count the number of dots in Braille letters (Braille-dot counting task), and evaluate the roughness of sandpaper discs (sandpaper task). Independently of the task, strong tactile evoked responses can be observed in occipital cortex after only 6 weeks after surgery (subject A1). The occipital cross-modal activation is much reduced after 15 weeks from surgery (subject A2). (B) BOLD responses elicited by visual stimuli (flashes of lights) before (blue) vs after (red) surgery in a group of RP patients implanted with Argus II. After the surgery visual BOLD signal is enhanced in V1. Crucially participants were scanned at least 6 months after the surgery. (C) RP patients were asked to choose the interval, demarcated by sound, in which a large high contrast grating was presented. The behavioral performance in a contrast detection task improves as a function of time from surgery. Patients needed time and practice to learn how to interpret and use the restored visual input. (A) Reproduced with permission from Figure 3 from Cunningham SI, Shi Y, Weiland JD, et al. Feasibility of structural and functional MRI acquisition with unpowered implants in Argus II retinal prosthesis patients: a case study. Trans Vis Sci Tech. 2015;4(6):6., ARVO copyright holder. (B, C) Reproduced from Figure 2 and Figure 5 from Castaldi et al. (2016) Visual BOLD Response in Late Blind Subjects with Argus e1002569. https://doi.org/10.1371/journal.pbio.1002569, published under the Creative Commons Attribution license (CC BY). Prosthesis. Biology II Retinal PLOS 14(10):
1701.05080
1
1701
2017-01-17T12:43:17
Individual versus collective cognition in social insects
[ "q-bio.NC", "cs.DC" ]
The concerted responses of eusocial insects to environmental stimuli are often referred to as collective cognition on the level of the colony.To achieve collective cognitiona group can draw on two different sources: individual cognitionand the connectivity between individuals.Computation in neural-networks, for example,is attributedmore tosophisticated communication schemes than to the complexity of individual neurons. The case of social insects, however, can be expected to differ. This is since individual insects are cognitively capable units that are often able to process information that is directly relevant at the level of the colony.Furthermore, involved communication patterns seem difficult to implement in a group of insects since these lack clear network structure.This review discusses links between the cognition of an individual insect and that of the colony. We provide examples for collective cognition whose sources span the full spectrum between amplification of individual insect cognition and emergent group-level processes.
q-bio.NC
q-bio
Individual versus collective cognition in social insects Ofer Feinermanᴥ, Amos Kormanˠ ᴥ Department of Physics of Complex Systems, Weizmann Institute of Science, 7610001, Rehovot, Israel. Email: [email protected] ˠ Institut de Recherche en Informatique Fondamentale (IRIF), CNRS and University Paris Diderot, 75013, Paris, France. Email: [email protected] Abstract The concerted responses of eusocial insects to environmental stimuli are often referred to as collective cognition on the level of the colony.To achieve collective cognitiona group can draw on two different sources: individual cognitionand the connectivity between individuals.Computation in neural-networks, for example,is attributedmore tosophisticated communication schemes than to the complexity of individual neurons. The case of social insects, however, can be expected to differ. This is since individual insects are cognitively capable units that are often able to process information that is directly relevant at the level of the colony.Furthermore, involved communication patterns seem difficult to implement in a group of insects since these lack clear network structure.This review discusses links between the cognition of an individual insect and that of the colony. We provide examples for collective cognition whose sources span the full spectrum between amplification of individual insect cognition and emergent group-level processes. Introduction The individuals that make up a social insect colony are so tightly knit that they are often regarded as a single super-organism(Wilson and Hölldobler, 2009). This point of view seems to go far beyond a simple metaphor(Gillooly et al., 2010)and encompasses aspects of the colony that are analogous to cell differentiation(Emerson, 1939), metabolic rates(Hou et al., 2010; Waters et al., 2010), nutrient regulation(Behmer, 2009),thermoregulation(Jones, 2004; Starks et al., 2000), gas exchange(King et al., 2015), and more. It is tempting to push this analogy, one step further and attribute the superorganism with collective cognition(Couzin, 2009; Franks, 1989; Seeley, 1996). In this respect, it is possible to envision two extreme cases in which groups of insects may have evolved to exhibit cognition on the scale of the entire colony. The first is reliance on the cognition of the individuals that make up the group. Indeed, the cognitive abilities of a single ant or bee within the large colony are far from being simple (Dornhaus and Franks, 2008). The group can benefit from these capabilities, for example, by sharing and refining the knowledge of informed individuals. The second extreme case is collective cognition derived from the interaction between members. Manmade systems teach us that complex computation can be achieved by the wiring together of very simple components such as logical gates (Lindgren and Nordahl, 1990). Similarly, social insect colonies often display dense interaction networks (Wilson and Hölldobler, 1988)and collective behaviors that appear to exceed the capacity of the individuals of which they are comprised (Sumpter, 2006). 1 It is therefore of interest to trace the collective actions of the social insect colony to their sources, be they the cognition of individuals or the communication network that connect groups of such individuals. Mapping out the relations between these two organizational scales is required if one is to understand and quantify collective cognition as well as learn about its evolutionary origins.We hypothesize that individual-based collective behaviors will be prevalent in cases where abilities, similar to those exhibited by solitary insects, suffice in order to sense, grasp, and process knowledge that is relevant on the scale of the colony. Deviations from this will tend to lead to group solutions that involve an increased emergent component. Outline The outline of this review is as follows: First, we discuss the cognitive abilities of the individuals that make up the social insect colony andsome current knowledge of communication networks in social insects. As mentioned above, these two components provide the basis on which the colony could build its collective capabilities. Next, we present a list of examples of collective cognition. These examples are ordered by the degree to which collective behaviors rely on each of the two components, from individual-based to connectivity-based. The examples are split into three categories: Individual-based collective behaviors, collective behaviors that combine different individual perspectives, and, finally, collective behaviors that display higher levels of emergence. Each of these categories is divided into subcategories that further refine this division. Taken together, these examples span a broad spectrum of relationships between individual and collective cognition. In the final section we discuss the possible factors that may determine the degree of emergence in a particular collective behavior. Individual cognition A good starting point for discussing the origins of cognition in social insect colonies is the cognitive abilities of solitary insects. Insect brains have evolved hundreds of millions of years(Farris and Schulmeister, 2010; Ma et al., 2012)prior to the appearance of eusociality(Moreau, 2006). Despite the fact that their brains are relatively small(Chittka and Niven, 2009),solitary insects exhibit high cognitive skills that include large behavioral repertoires(Evans, 1966), complex forms of learning(Alloway, 1972; Blackiston et al., 2011), and include navigational skills that often exceed those of humans(Brower, 1996). These abilities aid the solitary insect, among other things, inforaging (O'Neill, 2001), finding or constructing shelters(Raw, 1972), confronting predators(Schmidt, 1990), and identifying appropriate mating partners(Dickson, 2008). in this discussion The next step is the transitions to eusociality which happened between100-150 million years ago(Brady et al., 2006; Engel et al., 2009). Eusociality is characterized by reproductive division of labor that drastically lowers the level of conflict between group members as they strive towards common goals (Crespi and Yanega, 1995). It is first important to state the evident fact that, contrary to cells in a tissue or neurons in the brain, insects within the colony superorganism maintain their individuality. They are able of autonomous motion and decision making.Further, the brains of individuals within a colony 2 bear high resemblance to those of solitary insects(Strausfeld, 1976). One may therefore ask how the cognitive capabilities a social insect compare to those of a solitary insect. To date, it is not clear if once grouped into large groups evolution may work to increase or decrease the cognitive complexity of individuals. On the one hand, it is known, mainly from vertebrate groups that the communication requirements of group living may work to increase brain complexity(Shultz and Dunbar, 2010). On the other hand, it has been suggested that, relying on collective processes may ease the energetically expensive(Aiello and Wheeler, 1995) maintenanceof brain tissue(Anderson and McShea, 2001; Feinerman and Traniello, 2015). Whatever the exact comparison between the brains of a social and a solitary insect, it is clear that thesocial insect is a cognitively capable individual. Individuals within the colony possess the capacity for large behavioral repertoires(Chittka and Niven, 2009), for weighing a large number of factors to reach individual decisions(Franks et al., 2003), andfor navigating over large distances(Gathmann and Tscharntke, 2002; Wehner, 2003). Importantly, these individual capabilities are relevant on the scale of the entire colony. One aspectthat clearly differentiates the social insect from its solitary counterpart is the capacity for communication. For example, eusocial insects display a huge diversification of cuticular pheromones(van Wilgenburg et al., 2011) used to convey multiple signals that are unique to colony life(Howard and Blomquist, 2005). Another famous example is the honeybee waggle dance(Von Frisc, 1950). While solitary insects may have the motivation to conceal a newly found item for personal consumption(Byrne et al., 2003), the bees have evolved an elaborate communication scheme which allows them to share this location.These and other interaction skills form the foundation of the insect society. In the next section we discuss some of the properties of the communication networks via which social insects coordinate their activities. Interaction Networks collection is especially important(Wilson Group living animals combine personal and social information when deciding upon their next action(Rieucau and Giraldeau, 2011). In eusocial insects – the social component of information and Hölldobler, 1988). Correspondingly, the modalities of communication and richness of cues and signals is greatly enhanced. Social insects use a variety of olfactory(Martin and Drijfhout, 2009; Morgan, 2009), tactile(Razin et al., 2013), visual, and vibrational(Delattre et al., 2015; Roces et al., 1993)messages as well as multi-model combinations of these (Ramsden et al., 2009) in their communication. Broadly speaking, these can be divided into several groups: Some messages require direct contact between individuals and can thus be considered as local in both space and time. Other signals are local in time but not space and are typically employed as alarm signals (e.g. highly volatile pheromones(Blum, 1969)). Yet another group are signals that are local in space but not in time. This group includes stigmergic, indirect communication between insects in which one individual modifies the environment and a second individual arriving at its modified surroundings(Theraulaz and Bonabeau, 1999). Mass recruitment pheromone trails (Jaffe and Howse, 1979)and nest construction without a blueprint (Franks and Deneubourg, 1997)are two impressive examples of stigmergy. Note that, for the case of pheromonal location at the same some later time reacts to 3 communication, the time scales that characterize pheromonal communication are evolvable as they depend on chemical evaporation times that,indeed, vary between species and tasks(Morgan, 2009; Witte et al., 2007)(Holldobler and Wilson, 1990). Quantifying the communication patterns requires descriptive frameworksfor the different interaction types as described above. Contact dependent interactions can be described as time-ordered (Blonder and Dornhaus, 2011)communication networks(Fewell, 2003; Moreau et al., 2011). It has been shown that, in a laboratory setting, the high mobility of social insectsdictates that, after a sufficient time window, interactions occur between practically all possible pairs and the network become highly connected (Mersch et al., 2013). Stigmergic communication has been described by using the language of statistical mechanics (Richardson et al., 2011), or by employing cellular-automata tools typically used to describe self-organization processes (Khuong et al., 2016). These interactions are, inherently, one-to- many signaling and have been shown to extend the connectivity induced by contact dependent communication (Richardson and Gorochowski, 2015)..Adding long range communication such as that involving alarm pheromones, we obtain a picture of a system in which, at least to first order and over long enough time-scales, interactions can be described as well mixed. In other words, over time an insect receives signals from any other insect in the colony. Not only are interactions mixed they are also, to a large extent anonymous. With a few exceptions (Mallon and Franks, 2000; Tibbetts, 2002), it is reasonable to assume that individuals do not recognize which of the hundreds to hundreds of thousands of other individuals they are currently interacting with. While the previous discussion seems to suggest that interactions are completely ergodic, it is important to stress that the social insect colony is, by no means, devoid of structure. For example, ant nests and bee hives are often concentrically arrangedsuch that young insects reside in the deep center while older individuals occupy progressively occupy areas that are closer to the boundaries or entrance (Beshers and Fewell, 2001). Even when structure is initially lacking, self-organization and amplification of noise can work to create spatio- temporal patterns over time (Richardson et al., 2011; Theraulaz et al., 2003). Moreover, ant (Tschinkel, 2004) and termite (Noirot and Darlington, 2000) nests exhibit complex structures of rooms and corridors and these further reflect on the spatial distribution of individuals within the nest. It has been shown that different individuals tend to occupy specific locations or chambers within the nest (Jandt and Dornhaus, 2009; Sendova-Franks and Franks, 1995) and that this reflects on the probability that they interact with other individuals in other parts of the nest (Mersch et al., 2013; Pinter-Wollman et al., 2011). Spatial locations therefore inducea network structure that is composed of relatively stable clusters. Hence, in a very broad sense, the communication patterns in a social insect colony can be viewed as residing between a well-mixed(on the more local scale) and a fixed (on the global, cluster, scale) network. Collective cognition Having described some of the basic "cognitive toolbox" available to the colony, we go on to discuss its collective scale behavioral products. In what follows, we presenta non- comprehensive insect colonies. list of examples for collective behaviors in social 4 Theexamples where chosen while focusing on the different possible gaps between the knowledge, actions, and capabilities of individuals and those of the entire group. They are ordered by the extent of this gap and divided between individual based collective cognition (small gaps) and emergent collective cognition that builds on the interaction between insects (large gaps). This division is, by no means, strict. Individual based collective cognition In a eusocial colony the genetic conflict between individuals in the group is minimal(Queller and Strassmann, 1998). This leads to an alignment of interests which implies that it is generally advantageous for informed individuals to share their knowledge with other group members. Utilizing this information is useful for the group as well. Since the number of informed individuals may be small one could expect that their actions be too weak to elicit any significanteffect or, alternatively, they be averaged out against opposing actions performed by other, losing this useful information, it may be profitable for the group to amplify the actions of these focal individuals. Anysuch amplification should be regulated to prevent runaway behavior in case of mistakes. Such mistakes could arise from the informed individuals themselves: they may hold only partial information or be plain wrong, or from communication: noisy interactions may distort the original message. less-informed, colony members. Instead of Next, we discuss several examples ofamplification circuits that make the products of individual cognitionavailable, effective, anduseful at the level of the group. Unconditional amplification The simplest example is the alarm response. When an individual ant senses danger she not only directly reacts to it but further emits a volatile alarm pheromone(Blum, 1969). This pheromone spreads around the ant eliciting similar responses from her neighboring nest- mates. This positive feedback circuit quickly spreads the danger signal to affect a large number of individuals(Jeanson and Deneubourg, 2009). This not only increases the group's surveillance of its environment (the "many eyes principle") but also allows it to take collective actions towards, for example, protection of the nest. Similar behaviors are displayed by termiteswhere chemical communication is accompanied by vibrationalsignaling (Delattre et al., 2015). These collective positive feedback circuits provide an informed individual that senses danger immediate and direct control over the actions of the group. In other words,the gap between individual and collective cognition is, practically, nonexistent.The group forsakes regulation and out-weights this crucial survival response over the possible price paid by false alarms. Conditioned amplification The mass recruitment foraging trail occurs as a single first ant locates a food source. This ant then uses her navigational skills to return to the nest while laying a pheromone trail that recruits others to the food such that foraging commences. While an emergent process may work to straighten the trail and make it shorter (see below) the trail still follows the qualitative solution as initial recruiter(Holldobler, 1971). then communicated by first discovered and the 5 the group"double-checks"the Importantly, in a large number of species, ants strengthen the initial trail only on the way back from the target food source and only if they independently found it to be profitable(Beckers et al., 1992a; Mailleux et al., 2003; Wilson, 1962). This regulationcan be considered as "delayed" since it occurs only after an initial positive response to the initial ant. This provides a mechanism by which target communicated by the initial ant before continuing to amplify her effect even further. Itallows the colony to reduce its response to ants that may have outdated information or are, for some reason, confused. Since pheromones are volatile and have a finite lifetime, their concentration along the trail depends on the rate at which they are enhanced. A trail which is not enhanced will eventually disappear. Thus, delayed regulation further supplies a mechanism for calibrating the level of activity on the recruitment trail (Simon and Hefetz, 1992), discontinuingit once the food source is exhausted(Wilson, 1962). It further allows the system to escape local minima by switching to foraging on more profitable food sources when such are identified(Beekman and Dussutour, 2007). Amplification with early regulation Desert ants typically forage alone and display only a rudimentary form of recruitment(Amor et al., 2010; Razin et al., 2013). The recruitment process occurs as ants that are informed about a food source outside the nest attempt to alert their nest-mates using imperfect communication. The interactions used are noisy in the sense that a recruitment interaction may be ignored or, conversely, a non-recruitment interaction may induce an ant to leave the nest. Therefore, amplification of the initial signal must be regulated so that recruitment occurs only when a food source is present and runaway behavior that results from mere interaction noise is avoided. It was shown that desert ants regulate recruitment early on in the process, at the entrance chamber of their nest(Razin et al., 2013). This regulation is the result of two behavioral components. The firstis the fact that individuals "know that they know"(Greenwald et al., 2015). There is a clear difference in the way directly and non-directly informed insects react to interactions with others(Razin et al., 2013; Schultz et al., 2008; Stroeymeyt et al., 2011). For example, ants that have been to the food areasimply disregard interactions and maintain high walking speed to increase the number and effectiveness of their recruitment interactions(Razin et al., 2013). On the other hand, ants with second hand knowledge react to interactions in a more cautious manner. These ants can either upregulate or downregulate their propensity to be recruited depending on the state of the individual they interact with(Razin et al., 2013). These rules allow regulate collective behavior with minimaldependence on theirunreliablecommunication skills. Specifically, thisworks to decrease the chances that the actions of non-informed ants have global consequences on the state of the nest and leaves the stage, so to say, to the directly informed ants(Razin et al., 2013). The second regulatory component is an early negative feedback. All else being equal, non- informed ants tend to lower their propensity to exit with passing time. This leads to a collective threshold that dissipates the effects of random or isolated interactions such that their effect quickly dies away and this protects the system from noise. A persistent informed the ants to 6 ant with first-hand information can generate enough activity such that interaction rates increase(Gordon and Mehdiabadi, 1999) and the system moves over this recruitment threshold. This early feedback mechanism in which positive feedback occurs only if the system passes a set threshold is similar to the generation of spikes in neurons(Razin et al., 2013). Note that while delayed regulation (as described in the previous subsection) works to regulate the amplitude of the collective response and terminate it when the stimulus ends, early regulation works to prevent amplification in the first place. Combining individual perspectives Amplifying the optimal option The social insect colony may do more than amplifying individual decisions – it can, in fact, poll individuals to reach consensus choice regarding the best solution among several alternative options. Examples for this come from house-hunting(Visscher, 2007) behaviors in ants(Franks et al., 2002) and bees(Seeley et al., 2006). When assessing the quality of a potential new nest site scout ants have been shown to incorporate an intricate, individually based,evaluation scheme which combines the different attributes of this location (e.g. its volume, the size of the door, and the level of light) in a non-trivial way(Franks et al., 2003).There is evidence that this assessment results in a single grade given by the ant to the new location(Robinson et al., 2011; Robinson et al., 2014). The group does not make its own assessments of nest quality (a hypothetical example for this could be moving the whole colony to occupy the alternative nests one at a time and using the resulting ant density (Gordon et al., 1993)to accurately measure the area of each) but, rather, uses a quorum sensing as a polling mechanism to compare the assessments of its individual scouts(Franks et al., 2002; Seeley et al., 2006). With high probability, this leads to the colony choosing the best among the alternatives with the accuracy of the decision growing with the size of the group(Sasaki et al., 2013). House-hunting provides another fascinating example of how the action of the group may work to refine individual decisions: When comparing different nests with specific attributes individual ants are prone to violate theregularity principle of rational decision making(Sasaki and Pratt, 2011). This principle states that if option A is preferred over option B then this should not change upon introducing a third option, C, that is inferior to both. This fallacy is not specific to insects but, rather, affects many different animals including humans. However, when a whole colony is presented with the choice between these nests it will tend to make the rational choice(Edwards and Pratt, 2009; Sasaki and Pratt, 2011). This is because the polling often terminates before individual ants have had the chance to fail the regularity principle since this requires visiting multiple nests(Robinson et al., 2014). We provided several examples (many more exist) of how the actions and decisions of capable individuals reflect at the level of the group. The group does not create new solutions but ratherworks to amplify, average, poll, and refine the actions of individual members. This is done using collective communication circuits that involve positive and negative feedbacks and certain non-linearities. Amplification in dynamic settings 7 In dynamic, fast-evolving scenarios, Information can quickly become obsolete and the individuals that carry useful information (Robson and Traniello, 2002) change over time (Gelblum et al., 2015). This entails two inherent problems: First, the relevant individual at a specific point in time has to be identified by the group. Second, when an individual that is better updated appears the group must revert to following it instead. This scenario is realized during cooperative transport by longhorn crazy ants (Czaczkes and Ratnieks, 2013). When ants cooperatively transport a large food item they can often lose orientation and become unknowledgeable regarding the correct way to the nest. To correct their path, these ants rely on well-informed individuals that are in the vicinity of the load but unattached to it (Gelblum et al., 2015). Instead of identifying the ant that currently holds valuable navigational information, here again, the group relies on the "know that you know" principle. In other words, an informed ant acts in a manner that is different from the other carriers: She attaches to the object and, without heeding to others, pulls it in the direction she knows to be correct. At the same time the carrying ants "acknowledge that they don't know" and apply a different behavioral rule – which is, in a sense, pull in the direction in which the load is currently moving. Together, these different rules as applied by informed and non-informed ants allow the group to optimally amplify the force of the informed leader (Gelblum et al., 2015). Importantly, after a period of about 10 seconds, the newly attached leader loses her orientation. This former leader then adapts the behavioral rules of an ordinary carrier and may continue in this state for many minutes. The directionality of the carrying group is next corrected by the attachment of a new leader ant that happened to be informed at that particular time. The fast switching between leaders can be viewed as a mechanism that enables the group to escape being trapped at local minima in which the group displays coordinated motion but the direction is wrong. Collective bootstrapping of individual solutions Another form of colony level solutions that is based on the actions of a large number of individuals is trail shortening. Ants are famous for the ability to gradually decrease the length of their pheromone trail so that it finally draws a geodesic between the food source and the nest (Feynmann, 1985). This process can occur by the accumulated effect of ants that leave the trail (Deneubourg et al., 1983) and return to it a short distance away. Useful detours, i.e. those that "cut a corner" and slightly decrease the trail's length, are then amplified by the group while non-useful detours are abandoned (Deneubourg et al., 1983; Goss et al., 1989; Reid et al., 2011). Even though trail shortening utilizes segments that were offered by individual ants, it is inherently different from the amplification schemes described in the previous section. This difference is manifested in the fact that in the previous examples require that the informed ant "know that she knows" and assess the quality of the information that she holds. Conversely, during trail shortening ants that mark short-cuts are not required to hold any knowledge about the quality of their solutions. Rather, it is the group that either amplifies or 8 eventually ignores this alternative trail segments through a pheromone based positive- feedback mechanism. Emergent collective cognition So far, we focused on group level behaviors that gain their computational power by amplifying the actions and decisions of individuals. This differs from the notion of emergent cognition wherein collective scale processes allow the group to qualitatively transcend individual capabilities. An intuitive example for emergence is the different physical phases of matter. Here, minimal changes in temperature or the coupling between microscopic particles leads to qualitatively different macroscopic phases (e.g. the solid to liquid transition). It is therefore interesting to ask whether grouping together, not simple inanimate particles, but rather cognitive individuals with a memory and complex behavioral rules can be expected to display different and perhaps higher forms of emergence. Specifically, what forms of cognitive emergence occur in the case of social insects? In this section, we list several examples for emergent collective actions. As before, the examples are loosely ordered according to the increasing gaps between the individual and the group. Weighted response to multiple stimuli Since they are grouped into large ensembles individual insects are, inevitably, much smaller than the size of the colony and its territory. As a consequence, individuals cannot have direct access to large-scale environmental and internal colony conditions. Despite this, the colony as a whole must react to the full set of stimuli and appropriately divide the work force(Robinson, 1992).Models(Beshers and Fewell, 2001) suggestthat colony level division of labor can result from single insects with different task thresholds(Bonabeau, 1996)that resolve work demands which they locally experience(Franks and Tofts, 1994). Similar to the house-hunting example described above this colony level phenomenon relies on the cognitive assessments of individuals. The difference being that, in this case, the colony does not form a consensus around the solution of a single individual but rather divides the work force in a weighted manner according to information that is too spread out tobe available to any one individual. Division of labor can include more complex mechanisms, such as recruitment, that allow ants to employ not only personal but also social information in their decisions(Robinson et al., 2009a). Partial decoupling between individual and collective scales Cooperative transport is the process in which a group of ants retrieves a food item much too large for any of them to move on their own. During this process, the information available to individuals may be plainly misleading and counterproductive for the group's collective goals. This happens when trajectories to the nest as experienced by the small ants may be inaccessible to the large loads and even take it towards dead-ends that are difficult to escape. It was shown that to avoid such deadlocks, the macroscopic scale occasionally decouples from the possibly misleading information available at the microscopic scale(Fonio 9 et al., 2016). This mechanism allows the group to utilize beneficial information while using noise present at the group level to escape deadlocks (local minima) and avoid the potentially devastating consequences of fully relying on misleading information. Importantly, such decoupling does not require that any single individual detect at any point in time, whether information is valuable or misleading. Emergence in this system is evident as a separation between collective and individual behavior. This is evident, as the carrying group (and the food item they transport) does not follow a trajectory that was suggested by any single ant. Collective response independent of individual actions In some cases, the group can display effective reactions to stimuli that are not conceived by any individual. Such can be the case in which a cooperatively carrying group hits an obstacle. In such instances, instead of attempting to advance directly towards the nest, the group decouples from the actions of its individual members and goes into a perpendicular motion that takes the carried piece of food towards the edges of the obstacle(Gelblum et al., 2015). It was shown that this change in collective motion does not require that any individual ant be aware of the obstacle and individually change her behavior(Gelblum et al., 2016). Rather, the physical constraint induced by the obstacle directly affects the group as a whole. As a response, the group's mode of motion changes in a way that facilitates obstacle circumvention. Self-organization withouta blueprint A final example for emergent behavior involves nest construction. Social insects construct some of the most magnificent structures in the biological world(Theraulaz et al., 2003). This is done through a stigmergic process(Theraulaz and Bonabeau, 1999) in which individuals locally interact with features of the structure by adding (or removing, in the case of dug nests) building material to them(Franks and Deneubourg, 1997). This induces indirect communication as insects interact with the product of the action of their nest-mates. It was further shown that, in some cases, building materials are combined with a volatile construction pheromone(Khuong et al., 2016). This adds a temporal dimension to the physical structure and expands the possibilities for local rules and the complexity of the resulting structure(Khuong et al., 2016). In contrast to trail formation, for example, where the result of the collective effort is a refinement of almost complete structures suggested by individual ants, nest construction creates structures which appear to be far from the capabilities of any individual. Indeed, in this high form of emergence, it is unlikely that individuals have a blueprint of the desired final product. Nevertheless, they follow local rules and such that their collective effort results in the construction of intricate nests. Discussion The first three sections of this discussion follow the structure of the previous example section. We raise and then discuss some hypotheses regarding the prevalence of collective- scale behaviors which strongly rely on the cognitive capabilities of individuals. 10 Individual-based collective cognition An important factor to note is that many of the immediate requirements of the social group coincide with those of the individuals that comprise this group. Like the solitary insect, the colony must also scan the environment for food (Gordon, 1995), locate shelters (Franks et al., 2002), transport food (Gelblum et al., 2015), confront predators (Lamon and Topoff, 1981; Monographs, 2016), and care for brood (Siveter et al., 2014; Wang et al., 2015). As noted above, the cognitive abilities of individuals in a colony do not seem to greatly differ from those of their solitary counterparts. As such the cognition of individuals within a colony hold direct advantages to the group as a whole.TheIndividual based collective cognitionsection,above, providesexampleswherein the actions of a single individual suffice for directing the entire colony. Generally speaking, amplifying of the behavior of an individual requires two colony level processes: First,the identification of the specific individual of interest and then the amplification of its behavior. For example, as mentioned above, the first ant to find a food source has the capacity to trace the complete path between the nest and the food(Cammaerts and Cammaerts, 1980; Hölldobler, 1976; Wilson, 1962). This ant makes itself identifiable by laying pheromone markings on the surface as it heads back to the nest(Beckers et al., 1992a). Amplification occurs as ants that follow this trail enforce it with further pheromonal markings. Finally, mass foraging develops along the trail drawn out by the initial recruiting ant(Beckers et al., 1989). In many cases, solutions provided by different individuals are in conflict such that one solution is amplified this must come at the expense of others. In this case a third decision making process occurs alongside identification and amplification. Some mechanisms by which such collective decisions occur include the preferential reinforcement of preferred solutions, such as preferred food sources being marked by higher pheromone concentrations (Beckers et al., 1992b; Jaffe and Howse, 1979; Sumpter and Beekman, 2003)and cross inhibition between alternative emerging solutions, as occurs during honeybee recruitment (Nieh, 2010). Combining individual perspectives It is not always the case that an individual insect holds the complete solution to the colony's current needs. In fact, inherent constraints work to limit the value of individually held information. A first constraint involves the size of the individual when compared to that of its colony. Colonies reside over territories that are tens of meters (in the case of ants) or even kilometers (for wasps and bees) across and include large elaborate nest structures on a scale of several meters (Tschinkel, 2004). It is impossible for a single individual whose size is on the order of one centimeter to individually monitor these large areas. Other constraints are a single individual propensity to be badly informed, mistakenly wrong, or be limited by its cognitive capacity (Sasaki and Pratt, 2012). The second set of examples as above highlights behaviors in which the group holds useful information about the environment but this information is distributed among a large number of individuals. Using communication, these information fragments can be integrated 11 to yield collective decisions that take the "big picture" into account. Such integration can often be classified using two general schemes defined for animal groups in general: The "many eyes principle" (Ward et al., 2011) in which the group's capacity for surveillance increases with the number of alert animals and the "many wrongs principle"(Biro et al., 2006; Simons, 2004) in which averaging effects work to reduce "noise" at the scale of an individual animal to yield accurate collective action. Integration of information happens over a large number of contexts and species.Some recurring principles in this process include: distributed integration, nonlinearities (Sumpter and Beekman, 2003), positive and negative feedback loops (Franks et al., 2002; Nieh, 2010), use of interaction rates and cuing delays (Camazine, 1993; Greene et al., 2013), higher influence and lower propensity to be influenced by better informed individuals (Korman et al., 2014; Razin et al., 2013; Schultz et al., 2008), and correctly weighing individual vs. social information (Robinson et al., 2009b; Robinson et al., 2012). This section also presents examples in which colonies have evolved to achieve a collective task by amplifying the behaviors of individuals that work towards the small scale version of this same task. For example, in the context of cooperative transport the actions of a leader ant are exactly those she would take when individually transporting a small food item to the nest. The group allows these actions to be the driving force behind the transport of a large heavy item. Emergent group level cognition Last, some colony functions may fall outside of the solitary insect's behavioral or cognitive repertoire. These include large scale behaviors such as assessing the relation between the size of two objects when bothof these are much larger than the insect itself (Fonio et al., 2016; Gelblum et al., 2016), and active food dissemination (Camazine, 1993; Greenwald et al., 2015; Howard and Tschinek, 1981). In such cases, to achieve the group-level task, the relevant capabilitiesmust arise either by newly evolved individual traits or through a collective process that relies on the connectivity between individuals (a combination of these two options is also possible). Cases in which group level processes bestow the group with abilities that are qualitatively beyond those of its individual members can, by definition(De Wolf and Holvoet, 2005), be considered as a form of emergence. In the two examples of cooperative transport presented above, group problem solving capabilities increase beyond those of individuals. In these cases, the small size of individual insect may prevent it from grasping the relevant large-scale relationships between the carried load and the obstacle. Therefore, in these cases the problem solving on the level of the group must decouple to some extent from the array of, possibly misleading, solutions as offered by individuals, even if these are highly adept navigators. The mechanisms that enable such decoupling between the scales do not have to be complex. In these examples, group-level noise in scent mark following (Fonio et al., 2016) and the persistence that results from alignment of forces (Gelblum et al., 2016) suffice for efficient transport that interacts with the environment on the relevant scale of the ant team and the large carried load. 12 Stigmergic nest construction holds the potential for higher levels of emergence. Here, nest construction is carried out by individuals who follow simple local rules while constantly reacting to the environmental product of their previous actions (Grassé, 1959). Following these stigmergic principles (Theraulaz et al., 2003) may allow insects to construct elaborate nest architectures (Tschinkel, 2004) without any blueprint. The degree of emergence in such processes is difficult to define. On the one hand, by using a stigmergic process, a single insect, a solitary queen for example, may construct a structure for which it holds no internal representation (Camazine et al., 2001; Theraulaz and Bonabeau, 1999). A larger group in which each individual follows the same rules may achieve faster construction but the quality of emergence remains constant (Theraulaz and Bonabeau, 1999). On the other hand, it has been shown that pheromone deposition onto the constructed nest is an essential part of collective stigmergic nest construction. The addition of this time dependent component (pheromones evaporate) has the potential of allowing the group to achieve more complex forms of emergence not achievable by a single individual working alone (Khuong et al., 2016). Indeed, such processes can be modeled using the mathematics of cellular automata a theory which holds the potential of describing extreme forms of emergence(Wolfram, 2002). Prevalence of individual-based mechanisms Next, we raise somehypothesesin anattempt to provide rationale for the observed imminence of individuals to colony scale processes. As shown above, it is often the case that colony level behaviors are a consequence of the direct of amplification of individual actionsthat rely on individual cognition. This didn't have to be the case. One could imagine an evolutionary pathway where collective cognition is constructed from a network of different individuals each manifesting different, possibly more basic, capabilities.In this case, the actions of individuals don't directly coincide with the actions of the group or even with each other. Rather, group performance emerges from the coordination between individuals. An example for this are neurons in the brain. Neurons have evolved to be cells whose actions are purely computational such that their spiking activity bears no direct relations to the collective process at which they participate. Hypothetically, there is no a-priori prevention that such structures arise in a social insect colony to serve as the basis for highly efficient collective scale behaviors. Why, therefore, is this not the general case? We suggest that the first part of the answer to this question has to do with the availability and high quality performances that individual-based group level cognition can achieve. As stated above, by the time group cognition has evolved, individual insects were already developed independent organisms(Farris and Schulmeister, 2010; Ma et al., 2012). As such, they already possessed many of the cognitive resources that are required by the group. Furthermore, these individual capabilities are, in no sense, simple. For example, individual insects are highly adept navigators. To get from one place to another individual insects employ a toolbox consisting of multiple tactics, often applied in parallel. Across different insects, these include landmark navigation (Collett et al., 1993), dead reckoning (Collett and Collett, 2000), backtracking (Wystrach et al., 2013), and cognitive map (Gould, 1986)navigation by means of scents (Morgan, 2009), visual cues (Collett et al., 1993; Esch et 13 al., 2001; Wehner, 2003), temperature, and even magnetic fields. Collective cognition that relies on such individual capabilities holds the advantage of utilizing these highly non-trivial traits. The fact that individual insects possess high cognitive capabilities does not, however, suffice in explaining why these capabilities take a central role in many collective level processes.In the context of the navigation example as in the previous paragraph,it may very well be the case that a navigational toolbox that relies on the distributed actions of a large number of cooperating insects not only exists but, also, outperforms other, individual-based, schemes. Again, one may ask why this is not the general case and why the role of individuals has remained so pronounced throughout evolution. We hypothesize that the answer to this question has to do with the complexity of efficient distributed solutions given the inherent constraints that apply for a colony of autonomous individuals. We hypothesize that in many cases, simple distributed algorithms would not outperform what is readily achievable by amplifying individual cognition. We further hypothesize that distributed algorithms that do surpass simpler amplification processes may be expected to be highly complex and therefore difficult to evolve. Hence, the system may be trapped within local minima within the fitness landscape which utilize the individual cognitive components that have evolved at earlier stages. The next sectiondiscusses some of the non-trivial limitations on emergent behavioral solutions. Computational constraints on emergent collective cognition Emergence is often associated with the notion of a group that "exceeds the sum of its parts". Theoretically speaking, achieving such highly effective cooperation typically requires a substantial degree of coordination. An intuitive example for this is the parallel search problem in which a group of non-communicating random walkers that start at a given location aim to collectively cover an area and then share the profits of their findings. In this context, to "exceed the sum of its parts" implies that the group of N searchers cover the area more than N times faster than each individual, were it acting on its own. In fact, in many natural topologies (Alon et al., 2011; Efremenko and Reingold, 2009), the situation is very far from this. For example, in grid topologies when searchers do not communicate and as long as N is not too small, multiple random walks that start from a single point, typically achieve negligible speed-up in cover time when compared to a single searcher(Alon et al., 2011) (i.e., the group takes almost as much time to cover an area as a single one of its members). In this case, being part of a group becomes highly unbeneficial from an individual's point of view: food must be shared but without the advantage of obtaining it at a faster rate. Biological systems often achieve emergence using involved communication. The most celebrated example for this is that of the brain, viewed as a large ensemble of neurons. In the case of brains, such coordination heavily relies on the fact that the neurons are organized into stable networks such that the set of neighboring neurons of a given neuron remains, relatively, fixed. Synaptic plasticity allows a pair of neighboring neurons to undergo a mutual feedback process and fine-tune theirconnectivity(Hebb, 1949). This, as conceptually demonstrated by the Hopfield model (Hopfield, 1982), may provide the system 14 with its immense computational power. Indeed, Hopfield networks have been shown to be capable of universal computation in the Turing sense (Síma and Orponen, 2003). What can one expect for social insect colonies? On the one hand,these are very far from being non-communicating. On the other hand, interaction networks appear to be more loosely defined than those that characterize a brain. Indeed, in the insect colony fixed organization structures may be difficult to achieve for long periods of time, due to the inherent mobility and anonymity constraints. This lack of structure leads to a lack of knowledge that further constrains the system(Burgos and Polani, 2016). Namely, while learning in the brain happens by strengthening the synaptic connections between pairs of specific hard-wired (and therefore mutually identifiable) neurons, it is difficult imagine such mechanisms for freely moving, anonymous insects. Further, many brain functions are known to depend on precise interaction patterns that enable, for example, high levels of synchrony(Abeles, 2010). Such preciseness is difficult to imagine in the case of autonomous independent agents. It appears that things become even worse if communication and information flows across the systems are, themselves, limited or distorted (Feinerman and Korman, 2012; Feinerman et al., 2014; Korman et al., 2014; Razin et al., 2013). The highly dynamic environment and interchanging network structures that characterize the social insect colony make it difficult to implement error correcting mechanisms such as repeatedly sending the same message in orderto reduce distortion. As demonstrated in (Feinerman and Korman, 2012; Feinerman et al., 2014; Korman et al., 2014; Razin et al., 2013) such circumstances make even basic distributed tasks, such as rumor spreading, challenging. It is therefore reasonable to assume that emergent phenomena would be even more difficult to implement in such conditions. Summary Useful information exists at the level of the individual insect. This often suffices for the group's needs and indeed we have presented many examples in which the group follows individuals by amplifying their effect. In other cases, the inherent scale gap between individual and group may render information held by individuals partial, irrelevant, or even misleading to the collective goals of the group. In such cases, emergent collective phenomenon may kick in. Cognition that emerges at the level of the colony is subject to multiple constraints that mainly result from the fact that individuals maintain their autonomy as insects within the colony. Indeed, the most complex collective circuits described to date(Nieh, 2010; Pratt et al., 2005; Seeley et al., 2011) may be defined as simple when compared to the neural circuits that allow, for example, an ant to find her nest by using vector integration(Ofstad et al., 2011; Wehner, 2003). Interestingly, the ways in which collective cognition appears, even during a single behavior, are not mutually exclusive.An example for thiscomes from cooperative transport in Paratrechina longicornis ants. Indeed, during this behavior the ants simultaneously benefit from both individual based(Gelblum et al., 2015) and collective based cognition(Fonio et al., 15 2016; Gelblum et al., 2016). This balance between the organizational scales allows the system to enjoy the best of both the macroscopic and the microscopic worlds. Forming better connections between different collective behaviors and the levels of emergence that characterize them requires further research. We suggest that such research be focused on two avenues. The first is a computational study of the powers and limitations of natural distributed algorithms(Feinerman et al., 2014; Greenwald et al., 2015). Specifically, there is a need to better understand the complexity and evolvabilty of distributed solutions of different qualities. The second avenue involves empirical studies that attempt to trace the evolution of emergent, communication-based solutions. An example for this direction can be a comparative study of nest structures across a large number of phylogenetically related ant or termite species. Acknowledgements This work has received funding from the European Research Council (ERC) under the European Union's Horizon 2020 research and innovation program (grant agreement No 648032).O.F., incumbent of the Louis and Ida Rich career development chair, supported by Israel Science Foundation grant 833/15, would like to thank and the Clore Foundation for their ongoing support. A.K. We would like to than Yossi Yovel for his comments on this manuscript. 16 Bibliography Abeles, M. (2010). Synfire Chains. In Encyclopedia of Neuroscience, pp. 829–832. Aiello, L. C. and Wheeler, P. (1995). The Expensive-Tissue Hypothesis: The Brain and the Digestive System in Human and Primate Evolution. Curr. Anthropol.36, 199. Alloway, T. (1972). Learning and memory in insects. Annu. Rev. Entomol.17, 43–56. Alon, N., Avin, C., Koucky, M., Kozma, G., Lotker, Z. and Tuttle, M. R. (2011). Many Random Walks Are Faster Than One. Comb. Probab. Comput.20, 481–502. Amor, F., Ortega, P., Cerdá, X. and Boulay, R. (2010). Cooperative prey-retrieving in the ant Cataglyphis floricola: an unusual short-distance recruitment. Insectes Soc.57, 91–94. Anderson, C. and McShea, D. W. (2001). Individual versus social complexity, with particular reference to ant colonies. Biol. Rev. Camb. Philos. Soc.76, 211–37. Beckers, R., Goss, S., Deneubourg, J. L. and Pasteels, J. (1989). Colony size, communication, and ant foraging strategy. Psyche (Stuttg).96, 239–256. Beckers, R., Deneubourg, J.-L. and Goss, S. (1992a). Trail Laying Behavior During Food Recruitment In The Ant Lasius-Niger (L). Insectes Soc.39, 59–72. Beckers, R., Deneubourg, J. L. and Goss, S. (1992b). Trails and U-turns in the selection of a path by the ant Lasius niger. J. Theor. Biol.159, 397–415. Beekman, M. and Dussutour, A. (2007). How to Tell Your Mates Recruitment Mechanisms. 105–124. Behmer, S. T. (2009). Animal Behaviour: Feeding the Superorganism. Curr. Biol.19, R366– R368. Beshers, S. N. and Fewell, J. H. (2001). Models of division of labor in social insects. Annu. Rev. Entomol.46, 413–40. Biro, D., Sumpter, D. J. T., Meade, J. and Guilford, T. (2006). From compromise to leadership in pigeon homing. Curr. Biol.16, 2123–8. Blackiston, D., Briscoe, A. D. and Weiss, M. R. (2011). Color vision and learning in the monarch butterfly, Danaus plexippus (Nymphalidae). J. Exp. Biol.214, 509–520. Blonder, B. and Dornhaus, A. (2011). Time-ordered networks reveal limitations to information flow in ant colonies. PLoS One6, e20298. Blum, M. S. (1969). Alarm pheromones. Annu. Rev. Entomol.14, 57–80. Bonabeau, E. (1996). Quantitative study of the fixed threshold model for the regulation of division of labour in insect societies. Proc. R. Soc. B Biol. Sci.263, 1565–1569. Brady, S. G., Schultz, T. R., Fisher, B. L. and Ward, P. S. (2006). Evaluating alternative hypotheses for the early evolution and diversification of ants. Proc. Natl. Acad. Sci.103, 18172–18177. Brower, L. (1996). Monarch butterfly orientation: missing pieces of a magnificent puzzle. J. Exp. Biol.199, 93–103. Burgos, A. C. and Polani, D. (2016). An Informational Study of the Evolution of Codes and of Emerging Concepts in Populations of Agents. Artif. Life22, 196–210. 17 Byrne, M., Dacke, M., Nordstrom, P., Scholtz, C. and Warrant, E. (2003). Visual cues used by ball-rolling dung beetles for orientation. J. Comp. Physiol. A Sensory, Neural, Behav. Physiol.189, 411–418. Camazine, S. (1993). The regulation of pollen foraging by honey bees: how foragers assess the colony's need for pollen. Behav. Ecol. Sociobiol.32, 265–272. Camazine, S., Deneubourg, J.-L., Franks, N. R., Sneyd, J., Theraulaz, G. and Bonabeau, E. (2001). Self-organisation in biological systems. Cammaerts, M.-C. and Cammaerts, R. (1980). Food recruitment strategies of the ants Myrmica sabuleti and Myrmica ruginodis. Behav. Processes5, 251–270. Chittka, L. and Niven, J. (2009). Are bigger brains better? Curr. Biol.19, R995–R1008. Collett, M. and Collett, T. S. (2000). How do insects use path integration for their navigation? Biol. Cybern.83, 245–59. Collett, T. S., Fry, S. N. and Wehner, R. (1993). Sequence learning by honeybees. J. Comp. Physiol. A172, 693–706. Couzin, I. D. (2009). Collective cognition in animal groups. Trends Cogn. Sci.13, 36–43. Crespi, B. J. and Yanega, D. (1995). The definition of eusociality. Behav. Ecol.6, 109–115. Czaczkes, T. J. and Ratnieks, F. L. W. (2013). Cooperative transport in ants ( Hymenoptera : Formicidae ) and elsewhere. Myrmecological News18, 1–11. De Wolf, T. and Holvoet, T. (2005). Emergence versus self-organisation: Different concepts but promising when combined. Lect. Notes Comput. Sci. (including Subser. Lect. Notes Artif. Intell. Lect. Notes Bioinformatics)3464 LNAI, 1–15. Delattre, O., Sillam-Dussès, D., Jandák, V., Brothánek, M., Rücker, K., Bourguignon, T., Vytisková, B., Cvačka, J., Jiříček, O. and Šobotník, J. (2015). Complex alarm strategy in the most basal termite species. Behav. Ecol. Sociobiol. 1945–1955. Deneubourg, J. L., Pasteels, J. M. and Verhaeghe, J. C. (1983). Probabilistic behaviour in ants: A strategy of errors? J. Theor. Biol.105, 259–271. Dickson, B. J. (2008). Wired for sex: the neurobiology of Drosophila mating decisions. Science322, 904–909. Dornhaus, A. and Franks, N. R. (2008). Individual and collective cognition in ants and other insects (Hymenoptera: Formicidae). Myrmecological News11, 215–226. Edwards, S. C. and Pratt, S. C. (2009). Rationality in collective decision-making by ant colonies. Proc. Biol. Sci.276, 3655–61. Efremenko, K. and Reingold, O. (2009). How Well Do Random Walks Parallelize? In Approximation, Randomization, and Combinatorial Optimization. Algorithms and Techniques, pp. 476–489. Emerson, A. E. (1939). Social Coordination and the Superorganism. Am. Midl. Nat.21, 182– 209. Engel, M. S., Grimaldi, D. a. and Krishna, K. (2009). Termites (Isoptera): Their Phylogeny, Classification, and Rise to Ecological Dominance. Am. Museum Novit.3650, 1–27. Esch, H. E., Zhang, S., Srinivasan, M. V and Tautz, J. (2001). Honeybee dances communicate distances measured by optic flow. Nature411, 581–583. 18 Evans, H. E. (1966). The Behavior Patterns of Solitary Wasps. Annu. Rev. Entomol.11, 123– 154. Farris, S. M. and Schulmeister, S. (2010). Parasitoidism, not sociality, is associated with the evolution of elaborate mushroom bodies in the brains of hymenopteran insects. Proc. Biol. Sci.278, 940–51. Feinerman, O. and Korman, A. (2012). Memory lower bounds for randomized collaborative search and implications for biology. In Proceedings of International Symposium on Distributed COmputing (DISC), pp. 61–75. Feinerman, O. and Traniello, J. F. A. (2015). Social complexity, diet, and brain evolution: modeling the effects of colony size, worker size, brain size, and foraging behavior on colony fitness in ants. Behav. Ecol. Sociobiol.70, 1063–1074. Feinerman, O., Haeupler, B. and Korman, A. (2014). Breathe before speaking: efficient information dissemination despite noisy, limited, and anonymous communication. In Proceedings of the 2014 ACM symposium on Principles of distributed computing - PODC '14, pp. 114–123. Fewell, J. H. (2003). Social insect networks. Science301, 1867–70. Feynmann, R. P. (1985). The Amateur Scientist. In Surely, you're joking Mr. Feynmann!, p. 79. Bantam Books. Fonio, E., Heyman, Y., Boczkowski, L., Gelblum, A., Kosowski, A., Korman, A. and Feinerman, O. (2016). A Locally-Blazed Ant Trail Achieves Efficient Collective Navigation Despite Limited Information. Rev. Franks, N. R. (1989). Army Ants: A Collective Intelligence. Am. Sci.77, 138–145. Franks, N. and Deneubourg, J. (1997). Self-organizing nest construction in ants: individual worker behaviour and the nest's dynamics. Anim. Behav.54, 779–96. Franks, N. R. and Tofts, C. (1994). Foraging for work : how tasks allocate workers. Anim. Behav.48, 470–472. Franks, N. R., Pratt, S. C., Mallon, E. B., Britton, N. F. and Sumpter, D. J. T. (2002). Information flow, opinion polling and collective intelligence in house-hunting social insects. Philos. Trans. R. Soc. Lond. B. Biol. Sci.357, 1567–83. Franks, N. R., Mallon, E. B., Bray, H. E., Hamilton, M. J. and Mischler, T. C. (2003). Strategies for choosing between alternatives with different attributes: exemplified by house- hunting ants. Anim. Behav.65, 215–223. Gathmann, A. and Tscharntke, T. (2002). Foraging ranges of solitary bees. J. Anim. Ecol.71, 757–764. Gelblum, A., Pinkoviezky, I., Fonio, E., Ghosh, A., Gov, N. and Feinerman, O. (2015). Ant groups optimally amplify the effect of transiently informed individuals. Nat. Commun.6, 1–9. Gelblum, A., Pinkoviezky, I., Fonio, E., Gov, N. S. and Feinerman, O. (2016). The Ant Pendulum: problem solving through an emergent oscillatory phase. Proc. Natl. Acad. Sci.Accepted f,. Gillooly, J. F., Hou, C. and Kaspari, M. (2010). Eusocial insects as superorganisms Insights from metabolic theory. Commun. Integr. Biol.3, 360–362. Gordon, D. M. (1995). The expandable network of ant exploration. Anim. Behav.50, 995– 19 1007. Gordon, D. M. and Mehdiabadi, N. J. (1999). Encounter rate and task allocation in harvester ants. Behav. Ecol. Sociobiol.45, 370–377. Gordon, D. M., Paul, R. E. and Thorpe, K. (1993). What is the function of encounter pattern in ant colonies. Anim. Behav.45, 1083–1100. Goss, S., Aron, S., Deneubourg, J. L. and Pasteels, J. M. (1989). Self-organized shortcuts in the Argentine ant. Naturwissenschaften76, 579–581. Gould, J. L. (1986). The locale map of honey bees: do insects have cognitive maps? Science232, 861–863. Grassé, P. P. (1959). La reconstruction du nid et les coordinations interindividuelles chez Bellicositermes natalensis et Cubitermes sp. la th??orie de la stigmergie: Essai d'interpr??tation du comportement des termites constructeurs. Insectes Soc.6, 41–80. Greene, M. J., Pinter-Wollman, N. and Gordon, D. M. (2013). Interactions with combined chemical cues inform harvester ant foragers' decisions to leave the nest in search of food. PLoS One8, e52219. Greenwald, E., Segre, E. and Feinerman, O. (2015). Ant trophallactic networks: simultaneous measurement of interaction patterns and food dissemination. Sci. Rep.5, 12496. Hebb, D. O. (1949). The organization of behavior: a neuropsychological theory. Holldobler, B. (1971). Recruitment Behavior in Camponotus socius (Hym. Formicidae). Z. Vgl. Physiol.75, 123–142. Holldobler, B. and Wilson, E. O. (1990). The Ants. Harvard University Press. Hölldobler, B. (1976). Recruitment behavior, home range orientation and territoriality in harvester ants, Pogonomyrmex. Behav. Ecol. Sociobiol.1, 3–44. Hou, C., Kaspari, M., Vander Zanden, H. B. and Gillooly, J. F. (2010). Energetic basis of colonial living in social insects. Proc. Natl. Acad. Sci. U. S. A.107, 3634–3638. Howard, R. W. and Blomquist, G. J. (2005). Ecological, behavioral, and biochemical aspects of insect hydrocarbons. Annu. Rev. Entomol.50, 371–393. Howard, D. F. and Tschinek, W. R. (1981). The flow of food in colonies of the fire ant, Solenopsis invicta: a multifactorial study. Physiol. Entomol.6, 297–306. Jaffe, K. and Howse, P. E. (1979). The mass recruitment system of the leaf cutting ant, Atta cephalotes (L.). Anim. Behav.27, 930–939. Jandt, J. M. and Dornhaus, A. (2009). Spatial organization and division of labour in the bumblebee Bombus impatiens. Anim. Behav.77, 641–651. Jeanson, R. and Deneubourg, J. L. (2009). Positive feedback, convergent collective patterns, and social transitions in arthropods. In Organization of Insect Societies - From genome to sociocomplexity, pp. 460–482. Jones, J. C. (2004). Honey Bee Nest Thermoregulation: Diversity Promotes Stability. Science (80-. ).305, 402–404. Khuong, A., Gautrais, J., Perna, A., Sbaï, C., Combe, M., Kuntz, P., Jost, C. and Theraulaz, G. information shape ant nest (2016). Stigmergic construction and topochemical architecture. Proc. Natl. Acad. Sci.113, 1303–1308. 20 King, H., Ocko, S. and Mahadevan, L. (2015). Termite mounds harness diurnal temperature oscillations for ventilation. Proc. Natl. Acad. Sci.112, 11589–11593. Korman, A., Greenwald, E. and Feinerman, O. (2014). Confidence Sharing: An Economic Strategy for Efficient Information Flows in Animal Groups. PLoS Comput. Biol.10, e1003862. Lamon, B. and Topoff, H. (1981). Avoiding predation by army ants: Defensive behaviours of three ant species of the genus Camponotus. Anim. Behav.29, 1070–1081. Ma, X., Hou, X., Edgecombe, G. D. and Strausfeld, N. J. (2012). Complex brain and optic lobes in an early Cambrian arthropod. Nature490, 258–261. Mailleux, A.-C., Deneubourg, J.-L. and Detrain, C. (2003). Regulation of ants' foraging to resource productivity. Proc. R. Soc. London. Ser. B Biol. Sci.270, 1609–1616. Mallon, E. B. and Franks, N. R. (2000). Ants estimate area using Buffon's needle. Proc. Biol. Sci.267, 765–70. Martin, S. and Drijfhout, F. (2009). A review of ant cuticular hydrocarbons. J. Chem. Ecol.35, 1151–61. Mersch, D. P., Crespi, A. and Keller, L. (2013). Tracking individuals shows spatial fidelity is a key regulator of ant social organization. Science340, 1090–3. Monographs, E. (2016). Colony Defense Strategies of the Honeybees in Thailand Author ( s ): Thomas D . Seeley , Robin Hadlock Seeley and Pongthep Akratanakul Published by : Wiley Stable URL : http://www.jstor.org/stable/2937344 REFERENCES Linked references are available on JSTO. 52, 43–63. Moreau, C. S. (2006). Phylogeny of the Ants: Diversification in the Age of Angiosperms. Science (80-. ).312, 101–104. Moreau, M., Arrufat, P., Latil, G. and Jeanson, R. (2011). Use of radio-tagging to map spatial organization and social interactions in insects. J. Exp. Biol.214, 17–21. Morgan, D. E. (2009). Trail pheromones of ants. Physiol. Entomol.34, 1–17. Nieh, J. C. (2010). A negative feedback signal that is triggered by peril curbs honey bee recruitment. Curr. Biol.20, 310–5. Noirot, C. and Darlington, J. P. E. C. (2000). Termite Nests: Architecture, Regulation, and Defense. In Termites: Evolution, Sociality, Symbiosis Ecology, pp. 121–139. Springer. O'Neill, K. (2001). Solitary Wasps: Behavior and Natural History. Ofstad, T. A., Zuker, C. S. and Reiser, M. B. (2011). Visual place learning in Drosophila melanogaster. Nature474, 204–7. Pinter-Wollman, N., Wollman, R., Guetz, A., Holmes, S. and Gordon, D. M. (2011). The effect of individual variation on the structure and function of interaction networks in harvester ants. J. R. Soc. Interface8, 1562–73. Pratt, S. C., Sumpter, D. J. T., Mallon, E. B. and Franks, N. R. (2005). An agent-based model of collective nest choice by the ant Temnothorax albipennis. Anim. Behav.70, 1023– 1036. Queller, D. C. and Strassmann, J. E. (1998). Kin Selection and Social Insects. Bioscience48, 165–175. Ramsden, E., Adams, J., Holldobler, B., Shepherd, J. D., North, R. D., Jackson, C. W., Howse, 21 P. E., Bash, E., Millor, J., Am??, J. M., et al. (2009). Orientation cues used by ants. Insectes Soc.37, 101–115. Raw, A. (1972). The biology of the solitary bee Osmia rufa (L.) (Megachilidae). Trans. R. Entomol. Soc. London124, 213–229. Razin, N., Eckmann, J. and Feinerman, O. (2013). Desert ants achieve reliable recruitment across noisy interactions. J. R. Soc. Interface10, 20130079. Reid, C. R., Sumpter, D. J. T. and Beekman, M. (2011). Optimisation in a natural system: Argentine ants solve the Towers of Hanoi. J. Exp. Biol.214, 50–8. Richardson, T. O. and Gorochowski, T. E. (2015). Beyond contact-based transmission networks : the role of spatial coincidence. J. R. Soc. Interface12, 20150705. Richardson, T. O., Christensen, K., Franks, N. R., Jensen, H. J. and Sendova-Franks, A. B. (2011). Ants in a labyrinth: A statistical mechanics approach to the division of labour. PLoS One6,. Rieucau, G. and Giraldeau, L.-A. (2011). Exploring the costs and benefits of social information use: an appraisal of current experimental evidence. Philos. Trans. R. Soc. Lond. B. Biol. Sci.366, 949–57. Robinson, G. E. (1992). Regulation of division of labor in insect societies. Annu. Rev. Entomol.37, 637–665. Robinson, E. J. H., Feinerman, O. and Franks, N. R. (2009a). Flexible task allocation and the organization of work in ants. Proc. Biol. Sci.276, 4373–80. Robinson, E. J. H., Richardson, T. O., Sendova-Franks, A. B., Feinerman, O. and Franks, N. R. (2009b). Radio tagging reveals the roles of corpulence, experience and social information in ant decision making. Behav. Ecol. Sociobiol.63, 627–636. Robinson, E. J. H., Franks, N. R., Ellis, S., Okuda, S. and Marshall, J. a R. (2011). A simple threshold rule is sufficient to explain sophisticated collective decision-making. PLoS One6, e19981. Robinson, E. J. H., Feinerman, O. and Franks, N. R. (2012). Experience, corpulence and decision making in ant foraging. J. Exp. Biol.215, 2653–9. Robinson, E. J. H., Feinerman, O. and Franks, N. R. (2014). How collective comparisons emerge without individual comparisons of the options. Proc. Biol. Sci.281,. Robson, S. K. and Traniello, J. F. (2002). Transient division of labor and behavioral specialization in the ant Formica schaufussi. Naturwissenschaften89, 128–131. Roces, F., Tautz, J. and Hölldobler, B. (1993). Stridulation in leaf-cutting ants - Short-range recruitment through plant-borne vibrations. Naturwissenschaften80, 521–524. Sasaki, T. and Pratt, S. C. (2011). Emergence of group rationality from irrational individuals. Behav. Ecol.22, 276–281. Sasaki, T. and Pratt, S. C. (2012). Groups have a larger cognitive capacity than individuals. Curr. Biol.22,. Sasaki, T., Granovskiy, B., Mann, R. P., Sumpter, D. J. T. and Pratt, S. C. (2013). Ant colonies outperform individuals when a sensory discrimination task is difficult but not when it is easy. Proc. Natl. Acad. Sci.110, 13769–13773. Schmidt, J. O. (1990). Insect defenses: adaptive mechanisms and strategies of prey and 22 predators. SUNY Press. Schultz, K. M., Passino, K. M. and Seeley, T. D. (2008). The mechanism of flight guidance in honeybee swarms: subtle guides or streaker bees? J. Exp. Biol.211, 3287–95. Seeley, T. D. (1996). The Wisdom of the Hive. Seeley, T. D., Visscher, P. K. and Passino, K. M. (2006). Group Decision Making in Honey Bee Swarms. Am. Sci.94, 220–229. Seeley, T. D., Visscher, P. K., Schlegel, T., Hogan, P. M., Franks, N. R. and Marshall, J. A. R. (2011). Stop Signals Provide Cross Inhibition in Collective Decision-Making by Honeybee Swarms. Science. Sendova-Franks, a. B. and Franks, N. R. (1995). Spatial relationships within nests of the antLeptothorax unifasciatus(Latr.) and their implications for the division of labour. Anim. Behav.50, 121–136. Shultz, S. and Dunbar, R. (2010). Encephalization is not a universal macroevolutionary phenomenon in mammals but is associated with sociality. Proc. Natl. Acad. Sci. U. S. A.107, 21582–21586. Simon, T. and Hefetz, A. (1992). Dynamics of mass recruitment and its regulation in Tapinoma simrothi. Biol. Evol. Soc. Insects 325–334. Simons, A. M. (2004). Many wrongs: the advantage of group navigation. Trends Ecol. Evol.19, 453–5. Siveter, D. J., Tanaka, G., Farrell, U. C., Martin, M. J., Siveter, D. J. and Briggs, D. E. G. (2014). Exceptionally preserved 450-million-year-old ordovician ostracods with brood care. Curr. Biol.24, 801–806. Starks, P. T., Blackie, C. A. and Thomas D Seeley, P. T. (2000). Fever in honeybee colonies. Naturwissenschaften87, 229–231. Strausfeld, N. J. (1976). Atlas of an insect brain. Springer Science & Business Media. Stroeymeyt, N., Franks, N. R. and Giurfa, M. (2011). Knowledgeable individuals lead collective decisions in ants. J. Exp. Biol.214, 3046–3054. Sumpter, D. J. T. (2006). The principles of collective animal behaviour. Philos. Trans. R. Soc. Lond. B. Biol. Sci.361, 5–22. Sumpter, D. J. T. and Beekman, M. (2003). From nonlinearity to optimality: pheromone trail foraging by ants. Anim. Behav.66, 273–280. Theraulaz, G. and Bonabeau, E. (1999). A Brief History of Stigmergy. Artif. Life5, 97–116. Theraulaz, G., Gautrais, J., Camazine, S. and Deneubourg, J.-L. (2003). The formation of spatial patterns in social insects: from simple behaviours to complex structures. Philos. Trans. A. Math. Phys. Eng. Sci.361, 1263–82. Tibbetts, E. a (2002). Visual signals of individual identity in the wasp Polistes fuscatus. Proc. Biol. Sci.269, 1423–1428. Tschinkel, W. R. (2004). The nest architecture of the Florida harvester ant, Pogonomyrmex badius. J. Insect Sci.4, 21. van Wilgenburg, E., Symonds, M. R. E. and Elgar, M. A. (2011). Evolution of cuticular hydrocarbon diversity in ants. J. Evol. Biol.24, 1188–1198. 23 Visscher, P. K. (2007). Group decision making in nest-site selection among social insects. Annu. Rev. Entomol.52, 255–75. Von Frisc, K. (1950). Bees: their vision, chemical senses, and language. Cornell University Press. Wang, B., Xia, F., Wappler, T., Simon, E., Zhang, H., Jarzembowski, E. A. and Szwedo, J. (2015). Brood care in a 100-million-year-old scale insect. Elife2015, 4–11. Ward, A. J. W., Herbert-read, J. E., Sumpter, D. J. T. and Krause, J. (2011). Fast and accurate decisions through collective vigilance in fish shoals. Proc. Natl. Acad. Sci.108, E27–E27. Waters, J. S., Holbrook, C. T., Fewell, J. H. and Harrison, J. F. (2010). Allometric scaling of metabolism, growth, and activity in whole colonies of the seed-harvester ant Pogonomyrmex californicus. Am. Nat.176, 501–10. Wehner, R. (2003). Desert ant navigation: how miniature brains solve complex tasks. J. Comp. Physiol. A. Neuroethol. Sens. Neural. Behav. Physiol.189, 579–88. Wilson, E. O. (1962). Chemical communication among workers of the fire ant Solenopsis saevissima (Fr. Smith) 1. The Organization of Mass-Foraging. Anim. Behav.10, 134–147. Wilson, E. O. and Hölldobler, B. (1988). Dense heterarchies and mass communication as the basis of organization in ant colonies. Trends Ecol. Evol. (Personal Ed.3, 65–8. Wilson, E. and Hölldobler, B. (2009). The superorganism: The Beauty, Elegance, and Strangeness of Insect Societies. W. W. Norton & Company. Witte, V., Attygalle, A. B. and Meinwald, J. (2007). Complex chemical communication in the Formicidae). (Hymenoptera: longicornis Latreille ant crazy Chemoecology17, 57–62. Paratrechina Wolfram, S. (2002). A New Kind Of Science. Wolfram Media. Wystrach, A., Schwarz, S., Baniel, A. and Cheng, K. (2013). Backtracking behaviour in lost ants: an additional strategy in their navigational toolkit. Proc. Biol. Sci.280, 20131677. 24
1907.13118
1
1907
2019-07-29T12:33:41
Assessment and manipulation of the computational capacity of in vitro neuronal networks through criticality in neuronal avalanches
[ "q-bio.NC" ]
In this work, we report the preliminary analysis of the electrophysiological behavior of in vitro neuronal networks to identify when the networks are in a critical state based on the size distribution of network-wide avalanches of activity. The results presented here demonstrate the importance of selecting appropriate parameters in the evaluation of the size distribution and indicate that it is possible to perturb networks showing highly synchronized---or supercritical---behavior into the critical state by increasing the level of inhibition in the network. The classification of critical versus non-critical networks is valuable in identifying networks that can be expected to perform well on computational tasks, as criticality is widely considered to be the state in which a system is best suited for computation. This type of analysis is expected to enable the identification of networks that are well-suited for computation and the classification of networks as perturbed or healthy. This study is part of a larger research project, the overarching aim of which is to develop computational models that are able to reproduce target behaviors observed in in vitro neuronal networks. These models will ultimately be used to aid in the realization of these behaviors in nanomagnet arrays to be used in novel computing hardwares.
q-bio.NC
q-bio
Assessment and manipulation of the computational capacity of in vitro neuronal networks through criticality in neuronal avalanches Kristine Heiney∗†§, Ola Huse Ramstad‡, Ioanna Sandvig‡, Axel Sandvig‡, and Stefano Nichele∗ †Department of Computer Science, Norwegian University of Science and Technology, Trondheim, Norway ∗Department of Computer Science, Oslo Metropolitan University, Oslo, Norway ‡Department of Neuromedicine and Movement Science, Norwegian University of Science and Technology, Trondheim, Norway §Email: [email protected] Abstract -- In this work, we report the preliminary analysis of the electrophysiological behavior of in vitro neuronal networks to identify when the networks are in a critical state based on the size distribution of network-wide avalanches of activity. The results presented here demonstrate the importance of selecting appropriate parameters in the evaluation of the size distribution and indicate that it is possible to perturb networks showing highly synchronized -- or supercritical -- behavior into the critical state by increasing the level of inhibition in the network. The classification of critical versus non-critical networks is valuable in identifying networks that can be expected to perform well on computational tasks, as criticality is widely considered to be the state in which a system is best suited for computation. This type of analysis is expected to enable the identification of networks that are well-suited for computation and the classification of networks as perturbed or healthy. This study is part of a larger research project, the overarching aim of which is to develop computational models that are able to reproduce target behaviors observed in in vitro neuronal networks. These models will ultimately be used to aid in the realization of these behaviors in nanomagnet arrays to be used in novel computing hardwares. I. INTRODUCTION Current computing technology is based on the von Neu- mann architecture, in which tasks are performed sequentially and control, processing, and memory are each allocated to structurally distinct components. With this architecture, con- ventional computers struggle to cope with the rising demand for data processing and storage. Furthermore, although recent advancements in machine learning technology have conferred great advantages to our data handling capabilities, processing continues to be performed on conventional hardware that has no inherent learning capabilities and thus requires huge amounts of training data, computational time, and computing power. To continue to fulfill the rapidly growing computing de- mands of the modern day, it will be necessary to develop novel physical computing architectures that are scalable, capable of learning, energy-efficient, and fault-tolerant. The use of self- organizing substrates showing an inherent capacity for infor- mation transmission, storage, and modification [1] would bring computation into the physical domain, enabling improved efficiency through the direct exploitation of material and physical processes for computation [2 -- 4]. Some key properties of self-organizing systems that make them well-suited for computational tasks include their lack of centralized control and their adaptive response to changes in their environment [5]. Such systems are composed of many autonomous units that interact with each other and the environment through a set of local rules to give rise to organized emergent behaviors at a macroscopic scale. This type of spontaneous pattern formation is fairly common in nature, and there has been recent interest in determining how to develop interaction rules to generate various desired emergent behaviors [6], including those geared toward computation. In addition, it has been demonstrated that self-organizing substrates, such as magnetic arrays and self-assembling molecules, can be used as com- putational reservoirs -- untrained dynamical systems composed of a collection of recurrently connected units -- by training a readout layer to map the output of the physical system to a target problem [7]. The brain is an excellent example of a self-organizing system; it shows a remarkable capacity for computation with very little energy consumption and no centralized control, and scientists and engineers have long looked to the structure and behavior of the brain for inspiration. Neurons grown in vitro self-organize into networks that show complex patterns of spiking activity, which can be analyzed to gain insight into the network's capacity for information storage and transmission. This behavior indicates that in vitro neuronal networks may serve as a suitable computational reservoir [8] and could also provide insights into the characteristics and dynamics desired for more engineerable substrates. The aim of the present research project is to construct computational models that are able to reproduce desired be- haviors observed in electrophysiological data recorded from engineered neuronal networks. These models will provide insight into the behavior of the neurons and enable us to reproduce it in other substrates. The computational capabilities of the models and different physical substrates developed from the models will be explored and their dynamics characterized. 9 1 0 2 l u J 9 2 ] . C N o i b - q [ 1 v 8 1 1 3 1 . 7 0 9 1 : v i X r a This work is part of a project entitled Self-Organizing Com- putational substRATES (SOCRATES) [9], which aims to take inspiration from the behavior of in vitro neuronal networks toward the development of novel self-organizing computing hardwares based in nanomagnetic substrates. In addition to providing an avenue for the development of novel computational hardwares, the developed models are also expected to provide insight into the functionality of neuronal networks in healthy and perturbed conditions, where typi- cal perturbations include chemical manipulation or electrical stimulation. The dynamics of perturbed neuronal networks will also be modeled using the developed framework, and the computational capabilities and dynamics of the resulting models will be characterized. On the basis of this modeling, strategies of interfacing with perturbed networks to recover their dynamics will be explored. The behavior of perturbed networks and their capacity for recovery will also provide insight into the robustness of the computational capabilities of engineered self-organizing substrates against analogous damage or perturbation. The remainder of this paper is organized as follows. Section II gives some background on the analysis method used in this work to assess the criticality of in vitro neuronal networks toward identifying networks that may be considered well- suited for computation. Section III describes the methods of preparing the dissociated primary cortical networks evaluated in this study, which were cultured on a microelectrode array (MEA) to enable observation of their electrical behavior, and the analysis methods applied to the electrophysiological data recorded from the networks. The results from the prelimi- nary analysis of the recorded data are presented in Section IV. These results include the observation of the course of maturation of the networks and their response to chemical perturbation to increase inhibition. The implications of the presented results are discussed in Section V. A brief overview of the long-term plan for this research project is then given in Section VI, and section VII concludes the paper. II. BACKGROUND: NEURONAL AVALANCHES It has been theorized that the brain self-organizes into a critical state to optimize its computational properties; the foundations and theorized functional benefits of this behavior have been reviewed in recent works [10, 11]. A system in the critical state rests at the boundary between two qualita- tively different types of behavior. In the subcritical phase, a system shows highly ordered behavior characterized as static or oscillating between very few distinct states, whereas in the supercritical or chaotic phase, the system shows highly unpredictable, essentially random behavior. Near the transition point between these two regimes, the system is poised to effectively respond to a wide range of inputs as well as store and transmit information, making it ideal in terms of the capacity a system has for computation [1]. To determine whether a neuronal network is in the crit- ical state, we turn to the scaling behavior of network-wide avalanches of activity. As first defined by Beggs and Plenz [12], a neuronal avalanche is any number of consecutive time bins in which at least one spike is recorded, bounded before and after by time bins containing no activity, as shown in Fig. 1b. In their study, Beggs and Plenz [12] demonstrated that the probability distributions of the size and duration of neuronal avalanches follow a power law, indicating that the propagation of activity in the cortex is in the critical state [13]. It has been further demonstrated that criticality is established by a balance between excitation and inhibition and that cortical networks at criticality show a larger dynamic range (sensitivity to a wider range of inputs), higher information capacity (number of output patterns generated in response to different inputs), and greater information transmission (information shared between two recording sites) than networks functioning outside of criticality [14, 15]. Studies on the spontaneous activity of dissociated cortical networks have indicated that these networks tend to self-organize into the critical state over the course of their maturation, though not all networks settle into this state and there is disagreement on when the networks reach criticality [16 -- 18]. In the present study, an analysis method based on the size distribution of neuronal avalanches was applied to the analysis of in vitro primary cortical neuronal networks (Fig. 1a). This method is based on previous analysis performed on cortical networks (e.g., [16]). The aim of this method is to determine whether a given neuronal network is in the critical state, which is presumed to be beneficial for the network in terms of its capacity to store information and perform computation. the present As a long-term goal of research project, avalanche size distribution analysis is being applied to record- ings obtained from different types of in vitro neuronal net- works (e.g., human induced pluripotent stem cell (iPSC)- derived dopaminergic networks, see [19]) to assess the criti- cality of the networks. Emerging network dynamics in healthy and perturbed conditions will be studied, characterized, and classified. On the basis of the preliminary results obtained in this work, we report the development of two cortical networks from day in vitro (DIV) 7 to DIV 51 and their response to chemical perturbation on DIV 51. To the authors' knowledge, this work represents the first time a network has been manipulated into the critical state through chemical perturbation. Furthermore, this type of anal- ysis has also not yet been applied to the characterization of the neuronal network dynamics in in vitro disease models, which is the focus of ongoing and future work. III. METHODS Neuronal networks were prepared using primary cortical neurons. This section presents the methods for preparing and recording from the neuronal networks, as well as the data analysis methods applied to the recorded data. A. Preparation and electrophysiology of neuronal networks The neuronal networks assessed here were prepared as follows. Primary rat cortical neurons (Thermo Fisher) were seeded on a feeder layer of human astrocytes (Gibco, Thermo 2 Fisher) at a density of approximately 1 × 105 neurons per MEA. The networks were left to mature for 7 DIV prior to recording. The spontaneous electrophysiological activity of the network was recorded using a 60-electrode MEA together with the corresponding in vitro recording system (MEA2100- System, Multi Channel Systems) and software (Multi Channel Experimenter, Multi Channel Systems). Recordings were taken every second day for 15 min. Culture feedings occurred at the start of each week immediately after recording. γ-Aminobutyric acid (GABA, Sigma Aldrich) was added to the cortical networks at rising concentrations to disrupt the excitation-to-inhibition ratio by increasing network inhibition. Prior to perturbation, a recording was taken to establish a baseline. GABA was then added directly to the culture media in microliter volumes, and recordings were taken immediately after this perturbation. Different concentrations (10 and 50 µM for Network A and 5, 10, and 25 µM for Network B) were chosen to provide increasing degrees of perturbation; lower concentrations were used for Network B because it showed lower levels of activity. The 15-min recordings were performed in succession with increasing concentrations. Following the final perturbation, 90% of the culture media was replaced to return the cultures to the baseline state. B. Avalanche analysis Avalanches were detected according to the method de- scribed by Beggs and Plenz [12]. Spikes were detected using a simple thresholding method [20] based on the standard deviation of the noise of the signal after applying a bandpass filter with a pass band of 300 Hz to 3 kHz1. The spike detection threshold was set to 6, 7, and 8 standard deviations below the median of the signal for the cortical networks, and the results were compared to assess the effect of the detection threshold on the classification results. The spikes were binned into time bins equal to the average inter-event interval (IEI), which is the time between events recorded across all electrodes, and avalanches were detected as any number of consecutive active time bins (bins containing at least one spike) bounded before and after by empty time bins (Fig. 1b). The size of an avalanche is defined as the number of electrodes that were active during the avalanche. A power law was then fitted to the avalanche size distribu- tion data. This power law takes the form P (s) ∝ s−α, (1) where s is the avalanche size, P (s) is the probability of an avalanche having size s, and α is the exponent of the fitted power law. The exponent has been reported to take a value of α = 1.5 for in vitro neuronal networks [12, 16]. The fitting was performed using two different fitting methods, nonlinear regression (NLR) and maximum likelihood estimation (MLE), and the results were compared. The fit was applied over the size range of smin = 2 to the maximum detected avalanche 1Code for spike detection is available at https://github.com/SocratesNFR/ MCSspikedetection. (a) (b) Fig. 1: (a) Microscope images of the primary cortical networks analyzed in this study. Network A (shown here at DIV 49) showed activity on all electrodes, whereas Network B (shown here at DIV 51) partially detached from the electrode area of the MEA chip, resulting in approximately 12 -- 15 electrodes of the 60 showing the vast majority of the recorded network activity. (b) Definition of a neuronal avalanche. Each dot represents a spike recorded by one of the electrodes (Ch. 1 -- 4). A time bin is active when it contains at least one spike and empty when there are no spikes. An avalanche is defined as a sequence of consecutive active time bins preceded and followed by empty bins, and the size is the number of electrodes active during the avalanche. 3 size smax, with a cap at smax = 59 electrodes. The goodness of fit was computed following Clauset et al. [21]. Synthetic datasets were generated from the fitted distribution, and their Kolmogorov -- Smirnov (KS) distances from the theoretical dis- tribution were compared to the empirical KS distance. The fitting was rejected if the fraction p of synthetic KS distances that were greater than the empirical KS distance was less than 0.1 (p < 0.1)2. IV. RESULTS The avalanche size distributions of different in vitro neu- ronal networks were observed as the networks matured, and the networks were classified as being critical or not critical at each recording time point based on the fitting results. In this preliminary work, no rigorous analysis was yet applied to classify networks as sub- or supercritical; rather, only the goodness of fit of the size distribution to a power law was evaluated to assess whether or not the network was in a critical state during each analyzed recording. Preliminary classification of non-critical cases as sub- or supercritical was performed by visual inspection alone, where subcritical be- havior is characterized by exponential decay and supercritical by a bimodal distribution. The primary cortical networks evaluated here (Networks A and B; see Fig. 1a) were observed as they matured from DIV 7 to DIV 51, and then the change in the activity in response to chemical perturbation by GABA on DIV 51 was evaluated. These networks showed a large amount of activity with high- amplitude spikes, providing richer results than iPSC-derived dopaminergic networks investigated in a previous work [19] and allowing for the comparison of the results across multiple detection thresholds (6, 7, and 8 standard deviations below the median) and fitting techniques (NLR and MLE), as described in Section III-B. This section will first present the observed course of matu- ration for the two cortical networks based on their avalanche size distributions and then discuss how the addition of GABA affected the criticality of the networks. In both cases, the avalanche size distributions and fitting results will be com- pared for the different detection thresholds and fitting methods considered in this work. A. Spontaneous activity during maturation Both networks showed an initial low-activity phase from DIV 7 to DIV 12, in which the mean firing rate averaged across all active electrodes was less than 0.1 s−1 and fewer than approximately 1000 neuronal avalanches were detected within the size range (2 -- 59 electrodes) considered for the power law fitting. In this time period, there were considered to be too few avalanches for the fitting results to be reliable. In all subsequent recordings of spontaneous activity (DIVs 14 -- 51), both the NLR and MLE fitting methods produced fits that did not meet the criterion (p > 0.1) for the networks to be classified as critical, indicating that neither of the networks 2Code for avalanche detection and goodness of fit evaluation is available at https://github.com/SocratesNFR/avalanche. 4 (a) DIV 21 (b) DIV 21 (c) DIV 44 (d) DIV 44 (e) DIV 37 (f) DIV 37 Fig. 2: Raster plots and size distributions for Network A. (a) DIV 21. Raster plots for detection thresholds of 6 (top) and 8 (bottom) standard deviations. The raster plots show 0.5- s intervals of high network activity with the intervening silent periods eliminated for the sake of visualization. (b) DIV 21. Size distributions for detection thresholds of 6 and 8 standard deviations. A power law function with an exponent of α = 1.5 is plotted for reference. (c,d) Same as (a,b) for DIV 44. (e,f) Same as (a,b) for DIV 37. entered the critical state during the observation period. All subsequent discussion on classifying the networks as sub- or supercritical over the course of their maturation is thus based solely on visual inspection; as stated previously, the implementation of a rigorous classification method in this regard remains as a task for future work. Figure 2 shows a selection of raster plots and avalanche size probability distributions for Network A at three different time points (DIVs 21, 44, and 37). The raster plots and size distributions are shown for detection thresholds of 6 and 8 standard deviations at each time point. The raster plots show the first three network bursts in each recording with the periods of relative silence between the bursts removed for the sake of visualization. The first time point shown (DIV 21; Fig. 2a and 2b) is representative of the type of activity seen from DIV 14 to 21. During this period, the networks appeared to be in a subcritical state, which is characterized by an exponential size distribution showing a rapid drop in probability as the avalanche size increases. As shown in the raster plots, this corresponds to loose synchrony and relatively low levels of activity. The second time point (DIV 44; Fig. 2c and 2d) is representative of the type of activity seen from DIV 23 to 51. During this period, the networks appeared to be in a supercritical state, which is characterized by a bimodal size distribution showing many very large avalanches and few avalanches of intermediate size. The raster plots demonstrate that this type of activity corresponds to tight synchronicity across most of the electrodes and high activity levels. The final time point shown in Fig. 2 (DIV 37; Fig. 2e and 2f) is shown to demonstrate the robustness of the classification method based on the goodness of fit measure p against false positives. The size distribution obtained for a detection threshold of 8 standard deviations appears to follow a power law, showing high linearity in log -- log space with a slope cor- responding to the expected value of α = 1.5 (MLE: α = 1.62, NLR: α = 1.53). However, in both cases, the criterion p > 0.1 was not met (MLE: p = 0.005, NLR: p = 0.0). The analysis results at a threshold of 6 standard deviations more clearly show the deviation of the size distribution from the linear fit, displaying the same type of bimodal distribution seen at DIV 44 (Fig. 2d). This supports the reliability of the goodness of fit measure. B. Perturbation to increase inhibition On DIV 51, both networks showed highly synchronized activity and avalanche size distributions with peaks at large avalanche sizes, behavior indicative of a supercritical state. Figure 3a and 3e shows raster plots of the activity of Networks A and B, respectively, on DIV 51 before chemical pertur- bation, which demonstrate their high level of synchronicity.3 These plots were obtained with a detection threshold of 6 standard deviations. The corresponding avalanche size distri- butions are shown in Fig. 3c and 3g, respectively. As stated previously, Network B detached from some of the electrodes (see Fig. 1a), causing approximately 12 -- 15 of the electrodes to show the vast majority of the activity, especially during later recordings. This is apparent in both the raster plot (Fig. 3e) and the size distribution (Fig. 3g); there is a small peak in the probability of avalanches of size 12, followed by a very sharp drop off toward larger avalanches. 3Note that the raster plot for Network A shows 1 min of activity, whereas that for Network B shows 2 min. Because Network B showed much less activity than Network A both before and after perturbation, the type of activity is better captured by showing a larger time window. (a) Network A before perturbation (b) Network A after perturbation (c) Network A before perturbation (d) Network A after perturbation (e) Network B before perturbation (f) Network B after perturbation (g) Network B before perturbation (h) Network B after perturbation Fig. 3: Raster plots and avalanche size distributions before and after chemical perturbation by GABA (DIV 51, threshold of 6 standard deviations). (a) Raster plot of Network A before perturbation. (b) Raster plot of Network A after perturbation. (c) Probability distribution of avalanche sizes for Network A before perturbation. A power law with an exponent of α = 1.5 is plotted for reference. (d) Probability distribution of avalanches sizes for Network A after perturbation. The fitting results obtained by NLR and MLE are plotted along with their α and p values. (e -- h) Same as (a -- d) for Network B. 5 TABLE I: Power law fitting results for the two networks after chemical perturbation on DIV 51. The results are given for different spike detection thresholds and fitting methods. Bold text indicates a good fit (p > 0.1). Italicized results indicate that the number of avalanches N was considered insufficient for reliable fitting results (N (cid:46) 1000). Network Threshold Fitting method A (50 µM GABA) B (25 µM GABA) 8 st. dev. 7 st. dev. 6 st. dev. 8 st. dev. 7 st. dev. 6 st. dev. NLR MLE NLR MLE NLR MLE NLR MLE NLR MLE NLR MLE α 1.32 1.54 1.42 1.54 1.40 1.54 1.95 1.67 1.72 1.69 1.60 1.64 p 0.28 0.02 0.72 0.00 0.14 0.00 0.54 0.48 0.16 0.06 0.05 0.31 Figure 3b and 3f shows the raster plots of Networks A and B, respectively, after the addition of GABA to increase inhibition. After the addition of GABA, the amount of activity decreased, and the high level of synchronicity was broken. In contrast to the behavior prior to perturbation, the perturbed networks showed much less periodicity in their activity, with different groupings of electrodes firing within close temporal proximity. This is reflected in the avalanche size distributions for the two networks shown in Fig. 3d, which both appear to follow a power law. The fitting results obtained using the two fitting methods are also given in the size distribution plots. For Network A, NLR yielded a good power-law fit, whereas MLE did not; for Network B, the converse was true. The effect of the detection threshold and fitting method was further evaluated by assessing the agreement among the fitting results for the same recording under different evaluation parameters. The fitting results for the two networks after per- turbation on DIV 51 are given in Table I, with cases yielding a good fit (indicating the network is in the critical state) reported in bold. In cases where the number N of detected avalanches was insufficient to produce a reliable fitting (N (cid:46) 1000), the results are shown in italics. For Network A, the NLR fitting results indicate the network is in the critical state, with reliable results produced for thresholds of 6 and 7 standard deviations. For Network B, only a threshold of 6 standard deviations yielded enough activity for a reliable fitting, and the MLE results indicate the network is in the critical state. V. DISCUSSION The two cortical networks observed in this study never self- organized into the critical state; rather, they appeared to enter into a subcritical state after an initial low-activity phase and then settled into a stable highly synchronized supercritical state. Pasquale et al. [16] have demonstrated that not all dissociated cortical networks will reach the critical state during their normal course of maturation. They also reported that cultures tend to fall into a single preferred state (subcritical, supercritical, or critical) by around DIV 21 and tend not 6 to diverge from that state. However, other previous studies have demonstrated different courses of maturation for cortical networks. For example, Yada et al. [18] have indicated that cortical networks undergo early development from subcritical to supercritical before finally reaching a critical state after DIV 10. In contrast to this, Tetzlaff et al. [17] have described activity traversing from supercritical to subcritical and finally reaching criticality around DIV 58 (± 20 days). It is possible that the present networks may have reached criticality if left to mature further; however, it seems more likely that they would have remained in the supercritical state, as would be predicted by the type of trends observed by Pasquale et al. [16]. It will be necessary to perform more rigorous fittings of exponential curves to the size distribution to definitively demonstrate that the cultures showed subcritical behavior. This remains a task for future work. An important achievement in this work was the successful perturbation of the in vitro networks from supercritical to crit- ical. The networks showed increasingly synchronized activity accompanied by heightened activity levels as they matured from DIV 23 onward. A balanced excitation-to-inhibition ratio is widely considered to play an important role in enabling the emergence of critical behaviors, and previous studies have demonstrated the importance of this balance both through modeling (e.g., [16, 22]) and experimentally (e.g., [14, 15]). On the basis of this evidence, the networks were perturbed at DIV 51, when they showed highly synchronized supercritical behavior, by adding GABA; the aim of this perturbation was to artificially balance the excitation-to-inhibition ratio by increasing inhibition, thereby pushing the networks into the critical regime. The results indicate that the addition of GABA (50 µM to Network A and 25 µM to Network B) successfully broke the high level of synchronicity seen in the networks and brought them into the critical state (see Fig. 3). As stated in Section III, the GABA concentration was incrementally raised for each network (10 and 50 µM for Network A; 5, 10, and 25 µM for Network B). In both cases, the network only reached criticality at the highest considered molarity; however, in future work, smaller increments will be considered, as GABA also has the undesirable effect of reducing activity. The analysis of both the spontaneous and perturbed activity also demonstrated that it is important to consider different spike detection thresholds and fitting methods when conduct- ing the analysis. The preliminary results reported here show that the goodness of fit measure [21] was able to robustly identify non-critical behavior even when the size distribution showed apparent linearity in log -- log space and that lowering the threshold in this case supported the identification of supercritical activity (see Fig. 2f). When GABA was added, there was insufficient activity at certain thresholds to achieve a reliable fitting (see Table I), indicating the importance of an appropriate threshold. Furthermore, although the results appear to demonstrate critical behavior in Network B with a threshold of 6 standard deviations, it is possible that this threshold was too low and may have yielded false positives, as it is unlikely that there was actual activity on the electrodes where neurons had not adhered (see Fig. 1a). In the current experimental setup, a threshold of 7 standard deviations seems to have been most appropriate to balance these considerations. The results also indicate that more work is needed to determine whether NLR or MLE is best suited to yielding a good power law fit, as the two fittings did not produce consistent values of α or classification results. VI. PLAN FOR FUTURE RESEARCH This work represents a preliminary step in a larger research project, which will be described briefly here. The plan for this research project is divided into four stages. In the first stage, a data analysis framework will be developed, with the avalanche analysis method described here constituting a crucial part of this framework. The framework involves methods of extracting meaningful features from electrophysiological data recorded from in vitro neuronal networks. Such features include con- ventional parameters considered in electrophysiological data analysis, such as the mean firing rate, as well as more complex measures, such as entropy and measures of connectivity. The connectivity of the engineered networks will also be modeled using graph theory approaches. The avalanche method pre- sented here represents a useful tool for classifying networks as critical or non-critical. Other methods of classification and clustering of networks will also be explored. The second stage of the project involves the construction of computing models, such as cellular automata (CAs), random Boolean networks (RBNs), and recurrent neural networks (RNNs), that show behavior similar to that of the neuronal networks [23, 24]. The data analysis framework developed in the first phase will be used as a method of capturing the target behavior to be reproduced in the models, and this framework will be continually refined as we improve our understanding of the important aspects of neuronal behavior that contribute to their computational capabilities. These computing models are developed using evolutionary algorithms with appropriate fitness functions defined on the basis of the target behavior. Important features of the models, such as their input and output mappings and number of states, will be explored, and the dynamics of the models will be characterized. The third phase involves the use of the developed models and the in vitro neuronal networks as reservoirs to perform computational and classification tasks as a proof-of-concept using reservoir computing. The models from the second stage will be refined based on their performance as computing reservoirs. The final stage consists of the exploration of the application of the models developed in the second stage to the study of engineered neuronal networks under perturbed conditions mimicking pathologies related to the central nervous system (CNS). Networks that have had their synaptic function per- turbed will be modeled and analyzed using the developed methods to characterize how their behavior differs from that of unperturbed networks. Methods of interfacing with the per- turbed networks to restore their dynamics to the unperturbed state will then be explored. 7 VII. CONCLUSION The aim of this research project is to extract meaningful behaviors and features from electrophysiological data recorded from in vitro neuronal networks and construct models that reproduce these behaviors toward the eventual realization of novel computing substrates based in nanomagnetic materials. This paper reported the application of an avalanche size distribution analysis to electrophysiological data, representing a first step in the development of an analytical framework to extract target behaviors from such data. The results indicate that the avalanche analysis method applied here is able to successfully classify networks as critical or not critical and that chemical perturbation is a feasible method of inducing criticality when a network is in the supercritical state. In future work, more networks will be analyzed with the hopes of achieving self-organized criticality in some networks, and further investigations of the effects of chemical perturbation on both critical and supercritical networks will be conducted. it can be determined if a network is in a critical state, which gives an indication of its suitability for use in computational tasks. Furthermore, the results presented here indicate that it is possible to bring cortical networks from the supercritical state to criticality by increasing inhibition in the networks. This would be a useful approach to manipulating networks to make them viable for performing computational tasks. In addition to the computa- tional applications of this analysis, it is also expected to be useful in distinguishing healthy and perturbed networks and to provide insight into how different diseases affect neuronal connectivity and communication, which will be the target of future work. With this type of analysis, ACKNOWLEDGEMENTS This work was conducted as part of the SOCRATES project, which is partially funded by the Norwegian Research Coun- cil (NFR) through their IKTPLUSS research and innovation action on information and communication technologies under the project agreement 270961. REFERENCES [1] C. G. Langton, "Computation at the edge of chaos: Phase transitions and emergent computation," Physica D, vol. 40, pp. 12 -- 37, 1990. [2] H. Broersma, J. F. Miller, and S. Nichele, "Computational matter: Evolving computational functions in nanoscale materials," in Advances in Unconventional Computing. Springer, 2017, pp. 397 -- 428. [3] Z. Konkoli, S. Nichele, M. Dale, and S. Stepney, "Reser- voir computing with computational matter," in Compu- tational Matter. Springer, 2018, pp. 269 -- 293. [4] J. H. Jensen, E. Folven, and G. Tufte, "Computation in artificial spin ice," The 2018 Conference on Artificial Life: A Hybrid of the European Conference on Artificial Life (ECAL) and the International Conference on the Synthesis and Simulation of Living Systems (ALIFE), no. 30, pp. 15 -- 22, 2018. [5] F. Heylighen, "The science of self-organization and adap- tivity," in in: Knowledge Management, Organizational Intelligence and Learning, and Complexity, in: The En- cyclopedia of Life Support Systems, EOLSS. Publishers Co. Ltd, 1999, pp. 253 -- 280. [6] R. Doursat, H. Sayama, and O. Michel, "A review of mor- phogenetic engineering," Natural Computing, vol. 12, pp. 517 -- 535, 2013. [7] B. Schrauwen, D. Verstraeten, and J. Van Campenhout, "An overview of reservoir computing: Theory, applica- tions and implementations," in Proceedings of the 15th European Symposium on Artificial Neural Networks. p. 471-482 2007, 2007, pp. 471 -- 482. [8] P. Aaser, M. Knudsen, O. Huse Ramstad, R. van de Wijdeven, S. Nichele, I. Sandvig, G. Tufte, U. S. Bauer, Ø. Halaas, S. Hendseth et al., "Towards making a cyborg: A closed-loop reservoir-neuro system," in Proceedings of the European Conference on Artificial Life 2017. MIT Press, 2017. [9] (2018) SOCRATES: Self-organizing computational [Online]. Available: https://www.ntnu.edu/ substrates. socrates [10] W. L. Shew and D. Plenz, "The functional benefits of criticality in the cortex," The Neuroscientist, vol. 19, no. 1, pp. 88 -- 100, 2013. [11] J. Hesse and T. Gross, "Self-organized criticality as a fundamental propertie of neural systems," Frontiers in Systems Neuroscience, vol. 8, 2014. [12] J. M. Beggs and D. Plenz, "Neuronal avalanches in neo- cortical circuits," The Journal of Neuroscience, vol. 23, no. 35, pp. 11 167 -- 11 177, 2003. [13] P. Bak, C. Tang, and K. Wiesenfeld, "Self-organized criticality: An explanation of the 1/f noise," Physical Review Letters, vol. 59, pp. 381 -- 384, Jul 1987. [14] W. L. Shew, H. Yang, T. Petermann, R. Roy, and D. Plenz, "Neuronal avalanches imply maximum dy- namic range in cortical networks at criticality," Journal of Neuroscience, vol. 29, no. 49, pp. 15 595 -- 15 600, 2009. [15] W. L. Shew, H. Yang, S. Yu, R. Roy, and D. Plenz, "Information capacity and transmission are maximized in balanced cortical networks with neuronal avalanches," Journal of Neuroscience, vol. 31, no. 1, pp. 55 -- 63, 2011. [16] V. Pasquale, P. Massobrio, L. L. Bologna, M. Chiap- palone, and S. Martinoia, "Self-organization and neu- ronal avalanches in networks of dissociated cortical neu- rons," Neuroscience, vol. 153, pp. 1354 -- 1369, 2008. [17] C. Tetzlaff, S. Okujeni, U. Egert, W org otter, and M. Butz, "Self-organized criticality in developing neu- ronal networks," PLoS Computational Biology, vol. 6, no. 12, 2010. [18] Y. Yada, T. Mita, A. Sanada, R. Yano, D. J. Bakkum, A. Hierlemann, and H. Takahashi, "Development of neu- ral population activity toward self-organized criticality," Neuroscience, pp. 55 -- 65, 2017. [19] K. Heiney, V. Devold Valderhaug, I. Sand vig, A. Sand- vig, G. Tufte, H. Lewi Hammer, and S. Nichele, "Eval- uation of the criticality of in vitro neuronal networks: Toward an assessment of computational capacity," arXiv e-prints, p. arXiv:1907.02351, Jul 2019. [20] K. Heiney, J. Mateus, C. D. F. Lopes, E. Neto, M. Lamghari, and P. Aguiar, "µSpikeHunter: An advanced computational tool for the analysis of neuronal communication and action potential propagation in microfluidic platforms," Scientific Reports, vol. 9, no. 1, p. 5777, 2019. [Online]. Available: https: //doi.org/10.1038/s41598-019-42148-3 [21] A. Clauset, C. R. Shalizi, and M. E. J. Newman, "Power-law distributions in empirical data," SIAM Re- view, vol. 51, no. 4, pp. 661 -- 703, 2009. [22] P. Massobrio, V. Pasquale, and S. Martinoia, "Self- organized criticality in cortical assemblies occurs in con- current scale-free and small-world networks," Scientific Reports, 2015. [23] S. Nichele and A. Molund, "Deep learning with cel- lular automaton-based reservoir computing," Complex Systems, vol. 26, 2017. [24] S. Nichele and M. S. Gundersen, "Reservoir computing using nonuniform binary cellular automata," Complex Systems, vol. 26, 2017. 8
1803.03742
1
1803
2018-03-10T02:19:09
K-shell decomposition reveals hierarchical cortical organization of the human brain
[ "q-bio.NC" ]
In recent years numerous attempts to understand the human brain were undertaken from a network point of view. A network framework takes into account the relationships between the different parts of the system and enables to examine how global and complex functions might emerge from network topology. Previous work revealed that the human brain features 'small world' characteristics and that cortical hubs tend to interconnect among themselves. However, in order to fully understand the topological structure of hubs one needs to go beyond the properties of a specific hub and examine the various structural layers of the network. To address this topic further, we applied an analysis known in statistical physics and network theory as k-shell decomposition analysis. The analysis was applied on a human cortical network, derived from MRI\DSI data of six participants. Such analysis enables us to portray a detailed account of cortical connectivity focusing on different neighborhoods of interconnected layers across the cortex. Our findings reveal that the human cortex is highly connected and efficient, and unlike the internet network contains no isolated nodes. The cortical network is comprised of a nucleus alongside shells of increasing connectivity that formed one connected giant component. All these components were further categorized into three hierarchies in accordance with their connectivity profile, with each hierarchy reflecting different functional roles. Such a model may explain an efficient flow of information from the lowest hierarchy to the highest one, with each step enabling increased data integration. At the top, the highest hierarchy (the nucleus) serves as a global interconnected collective and demonstrates high correlation with consciousness related regions, suggesting that the nucleus might serve as a platform for consciousness to emerge.
q-bio.NC
q-bio
K-shell decomposition reveals hierarchical cortical organization of the human brain Nir Lahav*1, Baruch Ksherim*1, Eti Ben-Simon2,3, Adi Maron-Katz2,3, Reuven Cohen4, Shlomo Havlin1. 1. Dept. of Physics, Bar-Ilan University, Ramat Gan , Israel 2.Sackler Faculty of Medicine, Tel Aviv University, Tel-Aviv , Israel 3.Functional Brain Center, Wohl Institute for Advanced Imaging, Tel-Aviv Sourasky Medical Center, Tel- Aviv, Israel 4.Dept. of Mathematics, Bar-Ilan University, Ramat Gan, Israel *Authors contributed equally to this work Corresponding author: Nir Lahav, Dept. of Physics, Bar-Ilan University, Ramat Gan, Israel. E-mail: [email protected] Abstract In recent years numerous attempts to understand the human brain were undertaken from a network point of view. A network framework takes into account the relationships between the different ‎parts of the system and enables to examine how global and complex functions might emerge ‎from network topology. Previous work revealed that the human brain features 'small world' characteristics and that cortical hubs tend to interconnect among themselves. However, in order to fully understand the topological structure of hubs, and how their profile reflect the brain's global functional organization, one needs to go beyond the properties of a specific hub and examine the various structural layers that make up the network. To address this topic further, we applied an analysis known in statistical physics and network theory as k-shell decomposition analysis. The analysis was applied on a human cortical network, derived from MRI\DSI data of six participants. Such analysis enables us to portray a detailed account of cortical connectivity focusing on different neighborhoods of inter- connected layers across the cortex. Our findings reveal that the human cortex is highly connected and efficient, and unlike the internet network contains no isolated nodes. The cortical network is comprised of a nucleus alongside shells of increasing connectivity that formed one connected giant component, revealing the human brain's global functional organization. All these components were further categorized into three hierarchies in accordance with their connectivity profile, with each hierarchy reflecting different functional roles. Such a model may explain an efficient flow of information from the lowest hierarchy to the highest one, with each step enabling increased data integration. At the top, the highest hierarchy (the nucleus) serves as a global interconnected collective and demonstrates high correlation with consciousness related regions, suggesting that the nucleus might serve as a platform for consciousness to emerge. Key words: Network theory, K-shell decomposition, cortical hubs, Graph theory, functional hierarchies, consciousness "..And you ask yourself, where is my mind?" The pixies (Where is my mind) Introduction The human brain is one of the most complex systems in nature. In recent years numerous attempts to understand such complex systems were undertaken, in physics, from a network point of view (Carmi, 2007; Cohen and Havlin, 2010; Newman, 2003; Colizza and Vespignani, 2007; Goh et al., 2007). A network framework takes into account the relationships between the different ‎parts of the system and enables to examine how global and complex functions might emerge ‎from network topology.‎ Previous work revealed that the human brain features 'small world' characteristics (i.e. small average distance and large clustering coefficient associated with a large number of local structures (Achard et al., 2006; Bullmore and Sporns, 2009; He et al., 2007; Ponten et al., 2007; Reijneveld et al., 2007; Sporns et al., 2004; Sporns and Zwi, 2004; Stam et al., 2007; Stam and Reijneveld, 2007; van den Heuvel et al., 2008)), and that cortical hubs tend to interconnect and interact among themselves (Achard et al., 2006; Buckner et al., 2009; Eguiluz et al., 2005; van den Heuvel et al., 2008). For instance, van den Heuvel and Sporns demonstrated that hubs tend to be more densely connected among themselves than with nodes of lower degrees, creating a closed exclusive "rich club" (Collin et al., 2014; Harriger et al., 2012; van den Heuvel and Sporns, 2011; van den Heuvel et al., 2013). These studies, however, mainly focused on the individual degree (i.e. the number of edges that connect to a specific node) of a given node, not taking into account how their neighbors' connectivity profile might also influence their role or importance. In order to better understand the topological structure of hubs, their relationship with other nodes, and how their connectivity profile might reflect the brain's global functional organization, one needs to go beyond the properties of a specific hub and examine the various structural layers that make up the network. In order to explore the relations between network topology and its functional organization we applied a statistical physics analysis called k-shell decomposition (Adler, 1991; Alvarez-Hamelin et al., 2005a; Alvarez-Hamelin et al., 2005b; Carmi, 2007; Modha and Singh, 2010; Pittel et al., 1996; Garas et al., 2010) on a human cortical network derived from MRI and DSI data. Unlike regular degree analysis, k-shell decomposition does not only check a node's degree but also considers the degree of the nodes connected to it. The k-shell of a node reveals how central this node is in the network with respect to its neighbors, meaning that a higher k-value signifies a more central node belonging to a more connected ‎neighborhood in the network. By removing different degrees iteratively, the process enables to uncover the most connected area of the network (i.e., the nucleus) as well as the connectivity shells that surround it. Therefore, every shell defines a neighborhood of nodes with similar connectivity (see Fig. 1). A few studies have already applied this analysis in a preliminary way, focusing mainly on the network's nucleus and its relevance to known functional networks (Hagmann et al., 2008; van den Heuvel and Sporns, 2011). For instance, Hagmann et al. revealed that the nucleus of the human cortical network is mostly comprised of default mode network regions (Hagmann et al., 2008). However, when examined more carefully, k-shell decomposition analysis, as shown here, enables the creation of a topology model for the entire human cortex taking into account the nucleus as well as the different connectivity shells ultimately uncovering a reasonable picture of the global functional organization of the cortical network. Furthermore, using previously published k-shell analysis of internet network topology (Carmi, 2007) we were able to compare cortical network topology with other types of networks. We hypothesize that using k-shell decomposition would reveal that the human cortical network exhibits a hierarchical structure reflected by shells of higher connectivity, representing increasing levels of data processing and integration all the way up to the nucleus. We further assume that different groups of shells would reflect various cortical functions, with high order functions associated with higher shells. In this way we aim to connect the structural level with the functional level and to uncover how complex behaviors might emerge from the network. Figure 1: K-shell decomposition process. K-shell decomposition takes into account the degree of the node as well as the degree of the nodes connected to it. shows example This the difference of the K-shell method compared with regular degree count. Top panel: The whole network. The yellow node is a hub (k=5) and thus one might think that it would be in the nucleus. But on the first step of the process (k=1), two of its neighbors will be removed to the re- first (blue). When computing the degree of the remaining nodes we notice there are no more nodes with only one link. The remaining network is the 1st core. On step 2a (k=2), another two of its neighbors will be removed (black). Then, when re-computing the degree of each node (step 2b), the yellow node has a low degree (k=1) and will be removed to the second shell. The process stops in k=3 when the remaining nodes will be removed and no node will remain in is composed of remaining network in a given k step and the nucleus is defined as the final k- core in the process. The nucleus of this network is thus the 2nd- core, last remaining nodes (red). K-crust includes the nodes that have been removed until step k of the process. This network has 5 nodes in its 2nd-crust (blue, black and yellow. for more details see methods). the network. K-core shell the the group of the Materials and methods Imaging The networks for our analysis were derived from two combined brain imaging methods, MRI/DSI recorded by Patric Hagmann's group from University of Lausanne (for all the functions and data sets, please refer to : http://www.brain-connectivity-toolbox.net/). Using this data, clusters of gray matter formed the nodes while fibers of white matter formed the edges of the cortical network. In this technique, 998 cortical ROIs were used to construct the nodes of each network and 14,865 edges were derived from white matter fibers (for more specific details please see Hagmann P. et al. (Hagmann et al., 2008)). Six structural human cortical networks were transformed into six connection matrices by Patric Hagmann's group, derived from five right handed subjects (first two networks were derived from the same subject in different times). These connection matrices were utilized to calculate the network's properties and to apply the k-shell decomposition analysis. We used binary connection matrices ('1' –connected, '0' – disconnected) and not weighted connection matrices because of known difficulties in determining the appropriate weights and how to normalize them (Hagmann et al., 2007; Hagmann et al., 2003; Van Den Heuvel and Pol, 2010; van den Heuvel and Sporns, 2011). In order to connect between our structural network and known functional networks the 998 nodes were clustered into 66 known anatomical regions in accordance with Hagmann et al. (Hagmann et al., 2008). Network theory Several network characteristics were used in our analysis (see supplementary material 6 for further details): Degree (k) of a node is the number of edges that connect to the node. Hub is a node with degree above the average degree of the network. Distance between nodes is the shortest path between node i and node j. Average diameter (L) of the network is denoted by: 𝐿 = 1 𝑁(𝑁 − 1) ∑ 𝑑𝑖𝑗 𝑖≠𝑗 dij – distance between node i and node j ; N – total number of nodes in the network Local clustering coefficient (ci) of a node i reflects the probability that "my friend's friend will also be my friend" (computed for each node). Clustering coefficient is the average over all local ci and it provides estimation of the amount of local structures in the network. Topologically it means that the network will have a large quantity of triangles: Small-world networks are networks that are significantly more clustered than random networks, yet have approximately the same characteristic path length as random networks (high clustering coefficient and low average distance). Assortativity coefficient is the Pearson correlation coefficient of degree between pairs of linked nodes. Positive values indicate a correlation between nodes of similar degree, while negative values indicate relationships between nodes of different degree. Assortativity coefficient lies between −1 and 1. We also examined whether the cortical network exhibits a hierarchal structure (not to be confused with the hierarchies derived from k-shell decomposition analysis) in which hubs connect nodes which are otherwise not directly connected. Networks with a hierarchal structure have a power law clustering coefficient distribution- C~K-β which means that as the node degree increases (k) the clustering coefficient (C) decreases. The presence of hubs with low clustering coefficient means that the network has a hierarchal structure (since hubs connect nodes which are not directly connected, triangles with hubs are not frequent). Module structures: the network's modular structure (community structure), is revealed by subdividing the network into groups of nodes, with a maximally iicNC1 possible number of within group links, and a minimally possible number of between-group links. K-shell decomposition method In the k-shell decomposition method we revealed the network's nucleus as well as the shells that surround it. The k-shell of a node indicates the centrality of this node in the network with respect to its neighbors. The method is an iterative process, starting from degree k=1 and in every step raising the degree to remove nodes with lower or similar degree, until the network's nucleus is revealed, along the following steps: Step 1. Start with connectivity matrix M and degree k=1. Step 2. Remove all nodes with degree ≤ k, resulting in a pruned connectivity matrix M'. Step 3. From the remaining set of nodes, compute the degree of each node. If nodes have degree ≤ k, step 2 is repeated to obtain a new M'; otherwise, go back to step 1 with degree k=k+1 and M=M'. Stop when there are no more nodes in M' (M'=0). The k-shell is composed of all the new removed nodes (along with their edges) in a given k step. Accumulating the removed nodes of all previous steps (i.e. all previous k-shells) is termed the k-crust. The k-core is composed of the remaining network in a given k step and the nucleus is defined as the final k-core in the process. In the end of every step a new k-shell, k-crust and k-core are produced of the corresponding k degree. In the end of the process the nucleus is revealed with the most central nodes of the network, and the rest of the nodes are removed to the different shells (see Fig. 1). Typically, in the process of revealing the nucleus, all removed nodes in the k-crust eventually connect to each other forming one giant component. The uniqueness of k-shell decomposition method is that it takes into account both the degree of the node as well as the degree of the nodes connected to that node. Thus, we can examine groups of nodes, every group with its own unique connectivity pattern. In this way one can examine cortical anatomical regions according to their connectivity neighborhood. For each node in the network we determined its shell level (i.e. to which shell it belongs, or if it survived the whole process, it belongs to the highest level – the nucleus). We then calculated shell levels for every anatomical region, comprised of many nodes, according to the weighted average shell level of its nodes. ‎ Statistics and random networks In order to evaluate the significant of the properties of the cortical network each result was compared to that of a randomized network. The network was randomized by keeping the degree distribution and sequence of the matrix intact and only randomizing the edges between the nodes (Rubinov and Sporns, 2010). For each cortical network several random networks were computed with different amount of randomized edges (from 1% until 100% of the edges). This process was repeated several times iteratively. K-shell decomposition ‎was applied for each of the randomized networks. Since the results of the cortical network were resilient to small perturbations (1% of the edges randomized) we raise the amount of randomization. For greater amount of randomization the results were fixed around an average value after 5 iterations (or more) using 100% random edges. Thus we took the random networks to be with 100% randomized edges and 5 iterations. To assess statistical significance of our results across networks, permutation testing was used (Van Den Heuvel and Pol, 2010). Matrix Correlations across 6 networks were computed and compared with correlations obtained from 1000 random networks. These random network correlations yielded a null distribution comprised of correlations between any two networks obtained from the random topologies. Next, we tested whether the real correlations significantly exceeded the random correlations, validated by a p-value< 0.01. Moreover, the significance of the observed connectivity within and between hierarchies was evaluated using a random permutation test. In this test, each node was randomly assigned with a hierarchy, while preserving the connectivity structure of the graph as well as hierarchy sizes. This process was repeated 10,000 times, and in each repetition, the number of connections within each hierarchy and between each pair of hierarchies was recorded. For each pair of hierarchies, a connectivity p-value was calculated using the fraction of the permutations in which the number of connections linking them was equal or higher than this number in the real data. Resulting p-values were corrected for multiple comparisons using the false discovery rate (FDR) procedure thresholded at 0.05. Results Cortex network topology The results of the K-shell decomposition process revealed that the human cortex topology model has an "egg-like" shape (see Fig. 2). In the "middle", 22% (± 12%) of the networks' nodes formed the nucleus ("the yolk" in the egg analogy) and "surrounding" the nucleus about 77% (± 12%) of the removed nodes formed the shells. These removed nodes did not reach the nucleus and connected to each other to form one giant component. The nucleus has on average 217 nodes (± 117) and the giant component has on average 770 nodes (± 121). The rest of the nodes are isolated nodes. These removed nodes did not connect to the giant component, and essentially connect to the rest of the network solely through the nucleus (some nodes are not connect to any other node in the network and thus were removed; on average 9± 6 nodes per cortical network). Over all 6 networks, the average k-core of the nucleus was 19(±1), which means that during the iterative process the nucleus was revealed after the removal of 19(±1) shells. Thus, the minimum degree in the nucleus is 20 and the average degree of the nodes in the nucleus is 45 (±4). In comparison, the average degree across the entire cortical network is 29 (±1), demonstrating that the nucleus contains hubs with significantly higher degree than that of the average network. In addition, the nucleus had considerably lower average distance compared to the average distance of the entire cortical network (2 ±0.2 vs. 3±0.1 , respectively). This finding means that it takes 2 steps, on average, to get from one node to any other node in the 217 nodes of the nucleus. Fig. 2: Topology of the cortical network. Topology of the cortical network (middle) compared with the internet topology, after Carmi et al(Carmi, 2007) (left) and random cortex network (right). In the cortical network the nucleus consists of 20% of the nodes while the remaining 80% compose a one giant component from all the removed nodes in the different shells. Note, a much bigger nucleus in the random cortical network and contrary to the cortical network larger amount of isolated nodes in both random and internet topologies. The giant component is formed in a process similar to a first order phase transition with several critical points, as for the internet (Carmi, 2007; Pittel et al., 1996). In the beginning of the process islands of removed nodes were forming and growing, but at some stage all of these islands connect together to form the giant component (see Fig. S1 for more details). This abrupt phase transition occurred, on average, in k-crust 15 ±1 (i.e. big islands of removed nodes were formed in crust 14, comprised of all previous shells including shell 14, but in crust 15 all of these islands disappear and a single giant component is formed). There is no significant difference between the number of removed nodes that were added to crust 15 compared to crust 14, yet a phase transition had occurred, suggesting that the difference is in the amount of the removed hubs. In crust 15, for the first time, enough hubs (which connect to lower degree nodes) were removed at once and connect all the islands to form the giant component. Later, another critical point is observed. On average in crust 18 (±1), a very large amount of nodes are removed at once to join the giant component (on average 282 nodes comprising 37% ±10% of all the nodes in the giant component (see Fig. S1). This may suggest that the process reached yet another group of higher hubs which have been removed along with their connections. These hubs connect to significantly more nodes than the previous hubs leading to a massive removal of nodes. We also note that the giant component features small world characteristics similar to the entire network (C=0.4 for both giant component and the whole network, average distance is 3.6 ±0.5 for the giant component, slightly higher than that of the whole network (3±0.1), see Fig. S2). Cortex network topology in comparison to other networks The cortical network topology is found to be very different from the topologies of a randomized cortex or the internet network (at the autonomous systems level) which displayed a "medusa-like" shape (Carmi, 2007) (see Fig. 2). In addition to the nucleus and the giant component both random and internet topologies have a large amount of isolated nodes, forming the "medusa legs" in the medusa shape (on average 17% in the randomized cortical networks and 25% in the internet network, unlike close to 0.3% ±0.3% in the cortical network). In addition, the average nucleus size of the randomized cortex is nearly three times bigger than the average nucleus of the human cortex (56% vs. 20%). The cortical nucleus contains only 50% of the hubs, the rest fall on average in the last 4-5 shells before the nucleus, while in the random cortex 100% of all hubs reached the nucleus (see Fig. S3). A network that displays a significant amount of hubs on several levels and not just in the nucleus could support a hierarchical structure that enables modular integration, as evident in cortical function (Bassett et al., 2008; Christoff and Gabrieli, 2000; Gray et al., 2002; Northoff and Bermpohl, 2004; Northoff et al., 2006). Note that in the cortical network the hubs outside the nucleus start on average at shell 14-15 which supports the hypothesis that the first phase transition (shell 15±1 ) is due to the removal of those hubs (as mentioned above). Correlation between topology and known brain functions In the k-shell decomposition analysis the connections of a node as well as its neighborhood determine at which shell that node will be removed. Neighborhood of High degree will be removed in a higher shell, or might survive the entire process and be part of the nucleus. Therefore, the giant component is comprised of different shells which represent different neighborhood densities of connectivity. These shells, corresponding to known cortical networks, enable an effective examination of cortical hierarchical organization. We, therefore, examined the functional attributes of the nodes found in the nucleus and in all shells, by checking the shell level of every anatomical region (mapping how many nodes from the anatomical region have been removed to the different shells). Subsequently, we were able to score each anatomical region in accordance with its place in the network's hierarchy represented by its shell level. This characterization is demonstrated to be more accurate than just analyzing the average degree of each anatomical region (see Fig. S4 and supplementary material 1 for further details). Furthermore, we examined the nucleus and revealed known functional areas that are always found in the nucleus across all 6 networks (see Fig. 3). These areas comprise the entire bilateral midline region and overlap with five major functional networks: motor and motor planning, the default network, executive control network, high order visual areas and the salience network (see Table 1 for full details). In contrast, several known functional areas were never in the nucleus across all 6 networks. These areas include most of the right temporal lobe (e.g. the fusiform gyrus, A1, V5), right Broca and Wernicke homologues and right inferior parietal cortex. Interestingly, all the areas that never appear in the nucleus are from the right hemisphere. Furthermore, 70% of all the lowest shells are from the right hemisphere while 60% of the areas that are always in the nucleus belong to the left hemisphere (see supplementary material 2 for more details ‎). Fig. 3: Anatomical regions and the network nucleus. Brain maps displaying anatomical regions that are always in the nucleus (red) and never in the nucleus (blue). Note that all the regions that never reach the nucleus are from the right hemisphere. Next, we used the critical points that were observed during the giant component formation (see supplementary material 3 for more details) in order to detect and establish different hierarchies of shells. Briefly, the creation of the giant component corresponded to the shell threshold of a middle hierarchy and the creation of the nucleus corresponded to the threshold of a high hierarchy. This analysis resulted in three major hierarchal groups (low, middle and high) as portrayed in Figure 4. Fig. 4: Anatomical regions according to their hierarchies. Brain maps displaying cortical anatomical regions according to their hierarchies. Red – low hierarchy, green – middle hierarchy, blue – high hierarchy. One can divide the cortex to low hierarchy regions found in the lateral bottom part of the cortex, middle hierarchy in the lateral middle part of the cortex, and high hierarchy in the lateral top and midline part of the cortex. RH - Right hemisphere, LH - Left hemisphere The first hierarchal group consists of regions found in the lowest shells (average shell level 8.8, number of nodes/edges: 99/730 respectively). The removed nodes of this group are distributed across the shells with relatively high standard deviation (4.42, e.g. fusiform gyrus, entorhinal cortex, parahippocampal cortex. See Table 1 and Fig. 5 for full details). Notably, in this hierarchal group 75% of the regions were bilateral and 50% of the regions were never in the nucleus. The second hierarchy is a middle group which includes nodes found in the highest shells, but still not in the nucleus (number of nodes/edges: 335/4377 respectively). This group can be further subdivided to two subgroups, distributed middle and localized middle according to their average shell level and standard deviation. The average shell level of the distributed middle group is 14.5 (±3.07). This subgroup includes regions like right A1, right V5 and right Broca's homologue (for full details see table 1 and Fig. 5d). The average shell level of the localized middle group is 16.67 (±1.13). This subgroup includes regions like right wernicke homologue and right middle frontal gyrus. In the middle hierarchy 56% of the regions are bilateral and 40% of the regions are from the right hemisphere (in localized middle 88% right). 48% of the regions in this hierarchy were never found in the nucleus (for full details see Table 1 and Fig. 5c). The third group is the highest hierarchy which contains regions predominantly found in the nucleus (number of nodes/edges: 561/8430 respectively). This group can also be subdivided to two subgroups, distributed high and localized high according to their average shell level and standard deviation. Average shell level of distributed high is 16.92 ( ± 2.82). This subgroup includes the superior frontal gyrus, left Wernicke, left Broca and left V5. The average shell level of the localized high group is 19.30 (±0.97) and includes the precuneus and the cingulate cortex (for full details see Table 1 and Fig. 5). In this hierarchal group 69% of the regions were bilateral while 28% of the regions belonged to the left hemisphere. 44% of the regions in this hierarchy were always in the nucleus (66% in localized high). Altogether, all the regions that are always in the nucleus are from the high hierarchy while the regions that never reached the nucleus are from lower hierarchies. Fig. 5: Hierarchies of the cortical network. Top left panel: average shell level of the hierarchies. X-axis: hierarchy, Y-axis: shell level. Top right panel: an example of a single anatomical region representing each hierarchy (derived from average cortical network over all 6 networks. For exact data see supplementary data 1 and Fig. S5). Right Precuneus as an example of localized high hierarchy regions (blue). Notably, this area always reached the nucleus. Right caudal middle frontal as an example of localized middle hierarchy regions (green). Notably, this area never reached the nucleus. Right fusiform gyrus as an example of low hierarchy regions (red). Note the high standard deviation of the shell distribution. This region never reached the nucleus. X-axis: k-shell number; Y-axis: number of nodes. Dashed line: nucleus. Bottom left: Right lateral occipital cortex as an example of distributed middle hierarchy regions (striped green, localized middle hierarchy as above). X-axis: k-shell number; Y-axis: number of nodes. Dashed line: nucleus. Bottom right: Left precentral gyrus as an example of distributed high hierarchy regions (striped blue, localized high hierarchy as above). This area always reached the nucleus. X-axis: k-shell number; Y-axis: number of nodes, dashed line: nucleus. Using the shell score we could further estimate the average shell level of known functional regions or networks (see Table S2). Interestingly, average shell level often reflected known functional lateralization as detailed in Table 2. For instance, while Broca's area is found in the nucleus, its right homologues never reached the nucleus. In a similar way, Wernicke's area is found in the high hierarchy and its right homologue in the middle hierarchy, again never reaching the nucleus. Right primary motor region and right TPJ are found in the middle hierarchy (and also never reached the nucleus) whereas their left counterparts are found in the high hierarchy (and left primary motor region always reached the nucleus). The functional network with the highest average shell level was the default mode network (DMN) with a score of 18.1. 81% of its regions were found in the high hierarchy with 70% always reaching the nucleus. Following the DMN, the salience and the sensorimotor networks also demonstrate high average shell level (17.3 and 17.5, respectfully) reflecting their high functional relevance. These results are detailed in Table 2 and in supplementary material 2. Connections between hierarchies In order to examine the connections between the different hierarchies, we compared the number of connections within each hierarchy to the number of connections with other hierarchies (calculated as a percentage of its total connections). Within the lowest hierarchy it was found that only 22‎% ±6.33% were self-connections and the rest were distributed between the middle group (30% ‎±3.36%) and the highest group (48% ‎±4.24%). In the middle hierarchy approximately half of the connections (52% ‎±2.6%) were self-connections and 41.5% ‎±2.6% were linked to the highest group. Interestingly, only 7% ‎±0.77% of the connections from the middle hierarchy were linked to the lowest hierarchy. The highest hierarchy exhibited the highest levels of self- connections (72% ‎±1.6%). Only 22.5% ‎±1.5% of its connections were linked to the middle hierarchy and 6% ‎±0.6% to the lowest hierarchy (for more details see table S1). These findings suggest a flow of information from the lowest to the highest hierarchy with each step enabling greater local processing, possibly supporting increased data integration. We further tried to distinguish the differences between localized and distributed hierarchies. Distributed hierarchies have high standard deviation of the shell distribution and localized hierarchies have small standard deviation of the shell distribution (see fig. 5). Notably, while most of the edges of the localized hierarchies were mainly self-connections or connections to their distributed partner in the same hierarchy (e.g. distributed to localized middle), the distributed hierarchies displayed more connections to other hierarchies (~15% in distributed subgroups compared to only ~8% in localized subgroups) supporting their role in cross-hierarchy data integration. Moreover, many of these connections were also across similar categories (e.g. distribute middle with distribute high, app. 25%). Furthermore, the distributed and localized subgroups within the same hierarchy displayed a large amount of connections between themselves (~33% of their connections), supporting the fact that they originate from the same hierarchy. The significance of the observed connectivity within and between hierarchies was evaluated using a random permutation test. The results showed that connectivity within each hierarchy is significantly higher (FDR q<0.0005) and that connectivity between all hierarchies was significantly lower (FDR q<0.0005) than expected according the size of the hierarchies (see Figure 6). Fig. 6: Connections between hierarchies. The size of the hierarchies represents total amount of intra hierarchy connections. Connections within any hierarchy is found to be significantly higher (arrows) and connections between hierarchies was significantly smaller (dash arrows) than expected when taking into account the size of the ‎hierarchies, supporting the modularity nature of every hierarchy. Note the increased self-connections as the hierarchies increase (percent connections are normalized by the total amount of connections in each hierarchy). Discussion In the current study we applied the k-shell decomposition analysis to reveal the global functional organization of the human cortical network. Using this analysis we managed to build a model of cortex topology and connect the structural with the functional level. Our findings indicate that the human cortex is highly connected and efficient, compared to other networks, comprised of a nucleus and a giant component with virtually no isolated nodes. The giant component consists of different degree shells which represent different neighborhoods of connectivity, revealing the global properties of the cortical network. Together with the nucleus, these connectivity shells were categorized into three hierarchies representing an increasing number of regional connections, possibly supporting an increase in data processing and integration within each hierarchy. In accordance, the highest hierarchy was predominantly comprised of left and midline cortical regions (including regions of the default mode network) known to be associated with high-order functions (Northoff et al., 2006). Lastly, this collective of interconnected regions, integrating information throughout the cortex, might allow global properties such as consciousness to emerge. Network properties Our findings demonstrate, in accordance with previous work (Achard et al., 2006; Cohen and Havlin, 2010; Ekman et al., 2012) that the cortical network is resilient to small perturbations, highly organized, interconnected and much more efficient compared with a random cortical network or the internet network. K-shell decomposition analysis further proved to be more accurate and provide better resolution of network properties compared to standard methods (e.g. counting degrees, for full details see supplementary material 1). The two main components of the cortical network, the nucleus and the giant component, both have small world properties though they might serve different roles. A higher clustering coefficient of the giant component alongside short average distance of the nucleus suggest that the majority of local processing takes place within the giant component while the nucleus mainly adds shortcuts and global structures to the network. Indeed, although the nucleus is highly connected, it includes only 50% of all hubs unlike the random nucleus which includes all network hubs (see Fig. S2, S3). These 'peripheral' hubs were located in the giant component and, as previously suggested (Achard et al., 2006), might enable efficient data integration and local information processing. Hubs outside the nucleus might therefore, serve as local processors integrating information from lower shells and transfer it forward to a higher hierarchy, eventually reaching the nucleus (for more information see supplementary material 4). Network hierarchies and data integration K-shell decomposition analysis reveals that the creation of the giant component entails several critical points. From these critical points we could characterize three major neighborhoods of connectivity or three hierarchies (for more details see supplementary material 3). The regions in the lowest hierarchy appeared to be mostly involved in localized sensory perception (e.g. the fusiform face area and visual "what" stream (Goodale and Milner, 1992)). The different nodes within this hierarchy broadly distributed along the shells which might enable efficient data transfer and processing before sending it to higher hierarchies. The middle hierarchy is found to be composed of high shells with high degree nodes, though half of them never reached the nucleus, a property that separates these regions from the high hierarchy. Functional regions found in this hierarchy appeared to be involved in high cognitive functions and data integration. For instance, most of the auditory network and regions involved in the integration of audio and visual perception were found in the middle hierarchy. In addition 40% of the executive control network (including right dorsolateral PFC, a crucial region in executive control and working memory (Raz and Buhle, 2006)) and the dorsal visual stream (where\what stream (Goodale and Milner, 1992)) are found in this hierarchy. Broca's area was also located in the middle hierarchy as well as other homologue regions related to language such as Broca and Wernicke homologues. The high hierarchy contained regions predominantly found in the nucleus. All regions that reached the nucleus across all cortical networks are found in this hierarchy. Unlike other hierarchies, this unique hierarchy is a single, highly interconnected component, which enables high levels of data integration and processing, probably involved in the highest cognitive functions. In accordance, the high hierarchy exhibited the highest amount of self-connections across hierarchies suggesting that it processes data mostly within itself (see Fig. 6). The nucleus (represented by the high hierarchy) has a very strong overlap with the default mode network (81%), in accordance with the result of Hagmann et al (Hagmann et al., 2008), and also with the visual cortex (75%), sensorimotor network (75%) and salience network (71%). The visual dorsal stream and the executive control network also display 60% overlap with the nucleus. Interestingly, all the regions that never appear in the nucleus (across all 6 networks) belong to the right hemisphere, while a strong tendency towards the left hemisphere appeared when examining the nucleus. As mentioned above, all the regions that reached the nucleus are mostly midline or left hemisphere regions. Roughly speaking, the left hemisphere is comprised of high hierarchy regions and the right hemisphere is comprised of middle hierarchy regions (see Fig. 4 and supplementary material 3 and 2). Looking across hierarchies it's evident that the lowest hierarchy has the smallest amount of connections to other hierarchies and within itself; the middle hierarchy has more connections, almost equally distributed between itself and others; and the high hierarchy has the largest amount of connections, most of them within itself (see Fig. 6). Interestingly, self- connections within each hierarchy were significantly higher and between hierarchies significantly smaller, than expected in random control according the size of the ‎hierarchies. This finding suggests that every hierarchy can be seen as a different module mostly involved in self-processing and only then transfers information to other hierarchies (Bullmore and Sporns, 2009; Hagmann et al., 2008; van den Heuvel and Sporns, 2011). Regarding cross hierarchy connections, it is important to note that most of the connections between middle and high hierarchies occur in their distributed subgroups. This finding suggests that in every hierarchy distributed regions are more involved in data transfer and integration across hierarchies, while localized regions deal more with data processing. Assuming that data integration requires cross hierarchy connections (the amount of data that a hierarchy receives from other hierarchies – the centrality of the hierarchy (Rubinov and Sporns, 2010)) and data processing depend on interconnected regions (the amount of calculations taking place inside the hierarchy – specialized processing within densely interconnected module (Rubinov and Sporns, 2010)), then data integration and processing seem to increase as we step up in the hierarchies. These findings could therefore suggest a flow of information from the lowest to the highest hierarchy with every hierarchy integrating more data and executing further processing, in line with previous studies and theoretical work (Christoff and Gabrieli, 2000; Gray et al., 2002; Northoff et al., 2006; Damasio, 2000). The low hierarchy receives information, performs specific calculations with its small amount of intra connections and passes the information to the higher hierarchies. The middle hierarchy is further able to integrate more data and locally process more information. At the top, the nucleus receives the most information from all other hierarchies and executes further processing using its dense interconnections, suggesting its vital involvement in data integration within the cortical network. The Nucleus as a platform for consciousness The regions in the nucleus form one component and constitute the most connected neighborhoods in the cortical network with the highest degrees. In contrast to the giant component, which mostly exhibits local structures (i.e. high clustering coefficient), all the regions in the nucleus form global structures (see supplementary material 4) and densely connect within themselves creating a unique interconnected collective module all over the brain. The regions and profile of this collective module are consistent with previous work (Collin et al., 2014; Hagmann et al., 2008; van den Heuvel and Sporns, 2011), mostly comprised of posterior medial and parietal regions. Furthermore, in Hagmann et al's structural cortical core, 70% of the core's edges were self-connections, similar to our findings within the high hierarchy (72%). In addition, this structural core forms one module and connected with connector hubs to all other modules in the network, reflecting our results that the nucleus is a single interconnected module with increased global structures. These findings further suggest that the distributed high hierarchy is composed of such connector hubs, in charge of connecting other hierarchies with the nucleus. A strong inter-connected nucleus has also been demonstrated by Sporns et al suggesting a rich club organization of the human connectome (Collin et al., 2014; van den Heuvel and Sporns, 2011; van den Heuvel et al., 2013). Their results revealed a group of "12 strongly interconnected bihemispheric hub regions, comprising, in the cortex, the precuneus, superior frontal and superior parietal cortex". These six cortical regions were part of our more detailed interconnected nucleus which further includes more regions of the high hierarchy (see Table 1). This interconnected collective module creates one global structure, involving regions from all over the cortex, which may create one global function. Given recent theories that explain consciousness as a complex process of global data integration (Balduzzi and Tononi, 2008; Damasio, 2000; Dehaene and Naccache, 2001; Tononi and Edelman, 1998; Godwin et al., 2015), in particular Global Work space Theory and integrated information theory (Balduzzi and Tononi, 2008; Dehaene and Naccache, 2001; Tononi and Edelman, 1998), one can postulate that such global function could be related to conscious abilities. We therefore suggest that the global interconnected collective module of the nucleus can serve as a platform for consciousness to emerge. Indeed, all of the regions in the nucleus have been previously correlated to consciousness activities (Achard et al., 2006; Godwin et al., 2015; Goodale and Milner, 1992; Gray et al., 2002; Northoff and Bermpohl, 2004; Northoff et al., 2006; Christoff et al., 2009), especially midline and fronto-parietal regions. The nucleus, receiving the most information from all other hierarchies and integrating it to a unified global function, is therefore a perfect candidate to be the high integrative, global work space region in which consciousness can emerge (for more information see supplementary material 5). Study limitations Some limitation issues have to be taken into account when interpreting the current results. First, our network is limited only to the cortex; future studies should examine the entire brain network and its influence on the profile of the hierarchies or nucleus. It is possible, for instance, that regions within the low hierarchy (e.g. the fusiform gyrus) might belong to higher hierarchies and are affected by lack of subcortical regions (such as the hippocampus). Lastly, the structural connections of our network were mapped with DSI followed by computational tractography (Hagmann et al., 2008; Hagmann et al., 2007; Hagmann et al., 2003; Schmahmann et al., 2007). Although DSI has been shown to be especially sensitive with regard to detecting fiber crossings (Hagmann et al., 2008; Hagmann et al., 2007; Hagmann et al., 2003; Schmahmann et al., 2007), it must be noted that this method may be influenced by errors in fiber reconstruction, and systematic detection biases. Conclusions The current study used k-shell decomposition analysis in order to reveal the global functional organization of the human cortical network. Consequently, we built a model of human cortex topology and revealed the hierarchical structure of the cortical network. In addition, this analysis proved to be more accurate than standard methods in the characterization of cortical regions and hierarchies. Our findings indicate that the human cortex is highly connected and efficient, compared to other networks, comprised of a nucleus and a giant component with virtually no isolated nodes. The giant component consists of different connectivity shells, which we categorized into three hierarchies representing an increasing number of regional connections. Such a topological model could support an efficient flow of information from the lowest hierarchy to the highest one, with each step enabling more data integration and data processing. At the top, the highest hierarchy (the global interconnected collective module) receives information from all previous hierarchies, integrates it into one global function and thus might serve as a platform for consciousness to emerge. Acknowledgments We would like to thank Mr. Kobi Flax for his crucial role in data analysis, without him this work would not be accomplished! We would also like to thank Dr. Itay Hurvitz for his invaluable comments. We thank the European MULTIPLEX (EU-FET project 317532) project, the Israel Science Foundation, ONR and DTRA for financial support. Tables: Table 1: Cortical anatomical regions according to hierarchies Anatomical Region Localized High Side Function Paracentral lobule Mid SMA - sensorimotor network (Always) Caudal anterior cingulate cortex Caudal anterior cingulate cortex Inferior parietal cortex L R L Salience network (Always) Salience\Executive control network (Always) DMN, Sensorimotor network, Visual dorsal stream (Always) Posterior cingulate cortex Mid DMN (Always) Rostral anterior cingulate cortex Mid Salience\Executive control network, DMN (Always) Precuneus Mid Isthmus of the cingulate cortex Pericalcarine cortex Postcentral gyrus Superior parietal cortex Supramarginal gyrus Bank of the Superior temporal sulcus Cuneus Distributed high Superior frontal cortex Precentral gyrus R R L L L L R L L DMN (Always) DMN (Always) Primary visual area Primary somatosensory cortex - Sensorimotor network Executive control, Sensory integration, Sensorimotor network, Visual dorsal stream Wernicke area, TPJ Visual dorsal stream Visual DMN\ Executive\ Salience, Sensorimotor network (Always) Primary motor cortex - sensorimotor network (Always) Superior temporal cortex Pericalcarine cortex Pars orbitalis Middle temporal cortex Lateral occipital cortex Isthmus of the cingulate cortex Cuneus Rostral middle frontal cortex Superior parietal cortex Superior frontal cortex Postcentral gyrus Lingual gyrus Localized middle Inferior parietal cortex Caudal middle frontal cortex Bank of the superior temporal sulcus Supramarginal gyrus Superior temporal cortex Frontal pole Frontal pole Medial orbitofrontal cortex Distributed middle Pars triangularis Pars triangularis L L L L L L L L R R R R R R R R R R L R R L Wernicke ,TPJ, Visual dorsal stream Primary visual Executive control network V5 (Visual dorsal stream), DMN Primary visual, Visual ventral stream DMN Visual Executive control network, DMN Executive, sensory integration, Sensorimotor network, Visual dorsal stream DMN\ Executive\ Salience\ Sensorimotor network Primary somatosensory cortex - Sensorimotor network Visual DMN, Sensorimotor network, Visual dorsal stream (Never) Executive control network, Sensorimotor network (Never) Visual dorsal stream (Never) Wernicke homologue, TPJ (Never) Wernicke homologue, TPJ, Visual dorsal stream (Never) Executive control network Salience and executive control networks Stimulus-reward associations Broca homologue (Never) Broca Middle temporal cortex Pars opercularis Pars opercularis Inferior temporal cortex Inferior temporal cortex Rostral middle frontal cortex Pars orbitalis Transverse temporal cortex Temporal pole R R L R L R R R L V5 (Visual dorsal stream), DMN (Never) Broca homologue (Never) Broca Visual association, Visual ventral stream (Never) Visual association, Visual ventral stream Salience and executive control networks (Never) Salience and executive control networks (Never) Primary auditory cortex (Never) Salience network Lateral orbitofrontal cortex L+R Stimulus-reward associations Medial orbitofrontal cortex Precentral gyrus Caudal middle frontal cortex Lateral occipital cortex Low Temporal pole Parahippocampal cortex Parahippocampal cortex Fusiform gyrus Fusiform gyrus Entorhinal cortex Entorhinal cortex Lingual gyrus L R L R R R L R L R L L Stimulus-reward associations Primary motor cortex - Sensorimotor network Executive control network, DMN, Sensorimotor network Primary visual, Visual ventral stream Salience network (Never) Hippocampal support, Visual ventral stream (Never) Hippocampal support, Visual ventral stream Face recognition, Visual ventral stream (Never) Face recognition, Visual ventral stream Hippocampal support, Visual ventral stream (Never) Hippocampal support, Visual ventral stream Visual association DMN= Default mode network, TPJ=temporal parietal junction. Always= region that always reaches the nucleus for all networks, Never= region that never reaches the nucleus for all networks. Table 2: Laterality effects Anatomical region Left Precentral gyrus (primary motor cortex) High (always) Right Middle Inferior parietal High (always) Middle (never) Supra marginal gyrus (Wernicke area,TPJ) Superior temporal (Wernicke area ,TPJ) Lateral occipital cortex (primary visual) Lingual gyrus (visual association) Bank of the superior temporal sulcus (vision) Pars Orbitalis (executive control network) Middle temporal (V5, DMN) Rostral middle frontal cortex (executive control network, DMN) High High High Low High High High High Middle (never) Middle (never) Middle High Middle (never) Middle (never) Middle (never) Middle (never) Superior frontal cortex High (always) High Caudal middle frontal cortex (executive control network, DMN) Inferior temporal cortex (visual association) Pars triangularis (Broca homologue) Pars opercularis (Broca homologue) Temporal pole (salience network) Parahippocampal cortex Fusiform gyrus Entorhinal cortex Functional Networks Middle Middle Middle Middle Middle Low Low Low Middle (never) Middle (never) Middle (never) Middle (never) Low (never) Low (never) Low (never) Low (never) Dorsal stream (where stream) 100% high 80% middle (80% never) Ventral stream (what stream) 60% low 60% low (80% never) Auditory network 100% high 100% middle (100% never) Executive control network 77% high 55% middle Default mode network Salience network Sensorimotor network 89% high (55% always) 71% high (57% always) 60% (always) 40% (always) 83% (always) 17% (always) DMN= Default mode network, TPJ=temporal parietal junction. Always= region that always reaches the nucleus for all networks, Never= region that never reaches the nucleus for all networks. References: Achard S, Salvador R, Whitcher B, Suckling J and Bullmore E 2006 A resilient, low-frequency, small-world human brain functional network with highly connected association cortical hubs The Journal of neuroscience 26 63-72 Adler J 1991 Bootstrap percolation Physica A: Statistical Mechanics and its Applications 171 453-70 Alvarez-Hamelin J I, Dall'Asta L, Barrat A and Vespignani A 2005a k-core decomposition of Internet graphs: hierarchies, self-similarity and measurement biases arXiv preprint cs/0511007 Alvarez-Hamelin J I, Dall'Asta L, Barrat A and Vespignani A 2005b Large scale networks fingerprinting and visualization using the k-core decomposition. In: Advances in neural information processing systems, pp 41-50 Balduzzi D and Tononi G 2008 Integrated information in discrete dynamical systems: motivation and theoretical framework PLoS computational biology 4 e1000091 Bassett D S, Bullmore E, Verchinski B A, Mattay V S, Weinberger D R and Meyer-Lindenberg A 2008 Hierarchical organization of human cortical networks in health and schizophrenia The Journal of Neuroscience 28 9239-48 Buckner R L, Sepulcre J, Talukdar T, Krienen F M, Liu H, Hedden T, Andrews-Hanna J R, Sperling R A and Johnson K A 2009 Cortical hubs revealed by intrinsic functional connectivity: mapping, assessment of stability, and relation to Alzheimer's disease The Journal of Neuroscience 29 1860-73 Bullmore E and Sporns O 2009 Complex brain networks: graph theoretical analysis of structural and functional systems Nature Reviews Neuroscience 10 186-98 Carmi S 2007 A model of internet topology using K-shell decomposition PNAS 104 11150-4 Christoff K and Gabrieli J D 2000 The frontopolar cortex and human cognition: Evidence for a rostrocaudal hierarchical organization within the human prefrontal cortex Psychobiology 28 168-86 Christoff K, Gordon A M, Smallwood J, Smith R and Schooler J W 2009 Experience sampling during fMRI reveals default network and executive system contributions to mind wandering Proceedings of the National Academy of Sciences 106 8719-24 Cohen R and Havlin S 2010 Complex networks: structure, robustness and function: Cambridge University Press) Cohen R e a 2003 Scale free networks are ultra small Phys Rev Lett 90 Colizza V and Vespignani A 2007 Invasion threshold in heterogeneous metapopulation networks Physical Review Letters 99 148701 Collin G, Sporns O, Mandl R C and van den Heuvel M P 2014 Structural and functional aspects relating to cost and benefit of rich club organization in the human cerebral cortex Cerebral Cortex 24 2258-67 Damasio A R 2000 The feeling of what happens: Body and emotion in the making of consciousness: Random House) Dehaene S and Naccache L 2001 Towards a cognitive neuroscience of consciousness: basic evidence and a workspace framework Cognition 79 1-37 Eguiluz V M, Chialvo D R, Cecchi G A, Baliki M and Apkarian A V 2005 Scale-free brain functional networks Phys Rev Lett 94 018102 Ekman M, Derrfuss J, Tittgemeyer M and Fiebach C J 2012 Predicting errors from reconfiguration patterns in human brain networks Proceedings of the National Academy of Sciences 109 16714-9 Garas A, Argyrakis P, Rozenblat C, Tomassini M and Havlin S 2010 Worldwide spreading of economic crisis New journal of Physics 12 113043 Godwin D, Barry R L and Marois R 2015 Breakdown of the brain's functional network modularity with awareness Proceedings of the National Academy of Sciences 201414466 Goh K-I, Cusick M E, Valle D, Childs B, Vidal M and Barabási A-L 2007 The human disease network Proceedings of the National Academy of Sciences 104 8685-90 Goodale M A and Milner A D 1992 Separate visual pathways for perception and action Trends in neurosciences 15 20-5 Gray J R, Braver T S and Raichle M E 2002 Integration of emotion and cognition in the lateral prefrontal cortex Proceedings of the National Academy of Sciences 99 4115-20 Hagmann P, Cammoun L, Gigandet X, Meuli R, Honey C J, Wedeen V J and Sporns O 2008 Mapping the structural core of human cerebral cortex PLoS Biol 6 e159 Hagmann P, Kurant M, Gigandet X, Thiran P, Wedeen V J, Meuli R and Thiran J P 2007 Mapping human whole-brain structural networks with diffusion MRI PLoS ONE 2 e597 Hagmann P, Thiran J P, Jonasson L, Vandergheynst P, Clarke S, Maeder P and Meuli R 2003 DTI mapping of human brain connectivity: statistical fibre tracking and virtual dissection Neuroimage 19 545-54 Harriger L, Van Den Heuvel M P and Sporns O 2012 Rich club organization of macaque cerebral cortex and its role in network communication PloS one 7 e46497 He Y, Chen Z J and Evans A C 2007 Small-world anatomical networks in the human brain revealed by cortical thickness from MRI Cerebral cortex 17 2407-19 Modha D S and Singh R 2010 Network architecture of the long-distance pathways in the macaque brain Proceedings of the National Academy of Sciences 107 13485-90 Newman M E J 2003 The structure and function of complex networks SIAM review 45 167- 256 Northoff G and Bermpohl F 2004 Cortical midline structures and the self Trends in cognitive sciences 8 102-7 Northoff G, Heinzel A, de Greck M, Bermpohl F, Dobrowolny H and Panksepp J 2006 Self- referential processing in our brain-a meta-analysis of imaging studies on the self Neuroimage 31 440-57 Pittel B, Spencer J and Wormald N 1996 Sudden emergence of a giantk-core in a random graph Journal of Combinatorial Theory, Series B 67 111-51 Ponten S C, Bartolomei F and Stam C J 2007 Small-world networks and epilepsy: graph theoretical analysis of intracerebrally recorded mesial temporal lobe seizures Clin Neurophysiol 118 918-27 Raz A and Buhle J 2006 Typologies of attentional networks Nature Reviews Neuroscience 7 367-79 Reijneveld J C, Ponten S C, Berendse H W and Stam C J 2007 The application of graph theoretical analysis to complex networks in the brain Clin Neurophysiol 118 2317-31 Rubinov M and Sporns O 2010 Complex network measures of brain connectivity: uses and interpretations Neuroimage 52 1059-69 Schmahmann J D, Pandya D N, Wang R, Dai G, D'Arceuil H E, de Crespigny A J and Wedeen V J 2007 Association fibre pathways of the brain: parallel observations from diffusion spectrum imaging and autoradiography Brain 130 630-53 Sporns O, Chialvo D R, Kaiser M and Hilgetag C C 2004 Organization, development and function of complex brain networks Trends Cogn Sci 8 418-25 Sporns O and Zwi J D 2004 The small world of the cerebral cortex Neuroinformatics 2 145-62 Stam C J, Jones B F, Nolte G, Breakspear M and Scheltens P 2007 Small-world networks and functional connectivity in Alzheimer's disease Cereb Cortex 17 92-9 Stam C J and Reijneveld J C 2007 Graph theoretical analysis of complex networks in the brain Nonlinear Biomed Phys 1 3 Tononi G and Edelman G M 1998 Consciousness and complexity science 282 1846-51 Van Den Heuvel M P and Pol H E H 2010 Exploring the brain network: a review on resting- state fMRI functional connectivity European Neuropsychopharmacology 20 519-34 van den Heuvel M P and Sporns O 2011 Rich-club organization of the human connectome The Journal of neuroscience 31 15775-86 van den Heuvel M P, Sporns O, Collin G, Scheewe T, Mandl R C, Cahn W, Goñi J, Pol H E H and Kahn R S 2013 Abnormal rich club organization and functional brain dynamics in schizophrenia JAMA psychiatry 70 783-92 van den Heuvel M P, Stam C J, Boersma M and Pol H H 2008 Small-world and scale-free organization of voxel-based resting-state functional connectivity in the human brain Neuroimage 43 528-39
1812.00598
3
1812
2019-08-21T04:07:18
A Static Distributed-parameter Circuit Model Explains Electrical Stimulation on the Neuromuscular System
[ "q-bio.NC" ]
Finite Element Modeling (FEM) has been widely used to model the electric field distribution, to study the interaction between stimulation electrodes and neural tissue. However, due to the insufficient computational capability to represent neural tissue down to an atom-level, the existing FEM fails to model the real electric field that is perpendicular to neuron membrane to initiate an action potential. Thus, to reveal the real electrode-tissue interactions, we developed a circuit to model transmembrane voltage waveforms. Here, we show a distributed-parameter circuit model to systematically study how electrode-tissue interaction is affected by electrode position, input current waveform, and biological structures in the neuromuscular system. Our model explains and predicts various phenomena in neuromuscular stimulation, guides new stimulation electrode and method design, and more importantly, facilitates a fundamental understanding of the physical process during electrode-tissue interaction. In our model, myelin is assumed to be inductive. The voltage waveform resonance caused by this inductive myelin accounts for the much lower stimulation threshold to activate motoneurons than muscle fibers, which is observed with in vivo measurements. These findings confirmed the feasibility of studying electrode-tissue interaction using a proper distributed-parameter circuit. Our current application on the neuromuscular system also raises the possibility that this distributed-parameter circuit model could potentially be applied to study other neural tissues, including the Peripheral Nervous System (PNS) and the Central Nervous System (CNS).
q-bio.NC
q-bio
117576, Singapore 117456, Singapore Singapore Abstract The existing Finite Element Modeling (FEM) fails to model the real electric field that activates a neuron, because the computation capability is insufficient to represent the neural tissue down to an atom-level. Thus, to reveal the real electrode-tissue interactions, we adopt the idea of using a circuit to simulate voltage waveforms, as proposed in the Circuit-Probability theory (C-P theory). Here we show a distributed-parameter circuit model to systematically study how the interaction between electrode and tissue is affected by the electrode position, input current waveform, and the biological structures in the muscle. Our model explains and predicts various phenomena in muscle stimulation, guides new importantly, facilitates a fundamental stimulation electrode and method design, and more understanding of the physical process during electrode-tissue interaction. With some proper modifications, other neural tissues, including the Peripheral Nervous System (PNS) and the Central Nervous System (CNS), can also be studied with this distributed-parameter circuit model. A Static Distributed-parameter Circuit Model to Study Electrical Jiahui Wang1,2,3, Hao Wang1,2,3*, Xin Yuan Thow2, Nitish V. Thakor1,2, Shih-Cheng Stimulation on Muscle Tissue Yen1,2, Chengkuo Lee1,2,3* 1 Department of Electrical & Computer Engineering, National University of Singapore, 4 Engineering Drive 3, 2 Singapore Institute for Neurotechnology (SINAPSE), National University of Singapore, 28 Medical Drive, #05-COR, 3 Center for Intelligent Sensors and MEMS, National University of Singapore, 4 Engineering Drive 3, 117576, Electroceuticals, where electrical stimulation is delivered to the nervous system and muscles, are becoming widespread therapeutic solutions to people with neurological disabilities. For example, people with spinal cord injury (SCI) above the sixth cervical vertebra are unable to control extant limbs due to interruption to the motor pathway[1]. Functional electrical stimulation (FES) could benefit these people by restoring functional actions, like voluntary grasp, via the electrical stimulation of specific muscles. However, despite its application and medical promise, the electrophysiology of electrical stimulation is neither precise nor well understood. The history of electrophysiology dates back to 1770, when Luigi Galvani first discovered bioelectricity as he made a frog muscle twitch by accidentally creating a battery from surgical instruments[2]. The mechanism of electrical stimulation remained elusive until 1952, when Alan Hodgkin and Andrew Huxley proposed a quantitative description of membrane current on the unmyelinated squid giant axon[3]. In 1976, Donald McNeal first applied the established nerve axon models to explain excitation of nerve tissue, by bringing in the concept of shared electric field between the stimulation electrodes and the excitable nerve tissue[4]. With increased computing power, modern computational models further improved McNeal's method, by accounting for additional parameters, such as the anisotropic extracellular conductivity in electric field simulation and non-linear response of neuronal cells and axons. Such modern computational models include the field-neuron model, which has been applied to analyze electrical stimulation on the peripheral nervous system[5][6][7] and the central nervous system[8][9]. Despite the development of these computational models, there are large discrepancies between these computational models and experimental observations. Firstly, the tissue should not be considered as purely resistive[5-9], as recent indirect[10][11][12] and direct[13] evidence suggests that the extracellular medium consists of non-resistive components. Second, the deterministic gating property assumed in the computational models[5-9] contradicts the probabilistic gating property as observed in single ion channel recording[14-18]. In addition, there lacks a complete model that can systematically study how the interaction between the electrode and tissue is affected by the electrode position, input current waveforms, and the biological structures of the neural tissue. In this paper, we aim to achieve such a complete model that explains and predicts all various phenomena in muscle stimulation. In the circuit-probability theory (C-P theory) proposed in our previous submission[19], the composite biological structures in the tissue are neglected for circuit simplification. In this study, by considering motoneurons, muscle fibers and extracellular medium, the lumped-parameter circuit adopted in the C-P theory is now expanded to a distributed-parameter circuit, enabling a more detailed investigation of the interaction between stimulating electrode sites and excitable tissues. This muscle model can qualitatively explain many phenomena: 1. With different locations of the stimulation electrode sites relative to the target motoneuron (motoneuron-electrode position), the stimulation efficiency can either increase or decrease with respect to the spacing between the two electrode sites. 2. The motoneuron-electrode position determines the polarity of the cross-membrane voltage waveform, which is an unsolved issue using the lumped-parameter circuit in the C-P theory. 3. The force mapping curves measured with increasing current amplitude may form a certain shape with multiple curvatures, which has been observed in our previous experiments as well. A conventional explanation for this phenomenon is that multiple groups of motoneurons are sequentially recruited when the current amplitude is increased[20][21]. But in our model, this phenomenon can be theoretically derived and numerically calculated. 4. Instead of forming a force mapping curve (a measured curve of the generated force with respect to input current amplitude or pulse width) with multiple curvatures, we observe that the recruitment of multiple motoneuron groups will induce an unstable output force at the transition current, showing an abnormally high error bar in the force mapping curve. This phenomenon is reflected in the measurement data. It helps us to tell when an additional group of motoneurons is recruited during the experiment. Based on a comprehensive understanding of muscle stimulation, we propose an effective design of the stimulation electrode and a corresponding stimulation method to reduce muscle fatigue and control stimulation efficiency. Firstly, a double-side multiple-channel polyimide electrode design is proposed to reduce stimulation fatigue by alternatively activating the motoneurons on the frontside and backside of the electrode. Secondly, a comprehensive calibration of the force mapping curve using different combinations of the electrode sites is necessary because the stimulation efficiency and linearity of force control using a specific electrode pair is determined by the motoneuron-electrode location, meaning that a universal parametric force control method does not exist. In addition, we propose new evidence to support the idea of inductive myelin sheath on motoneuron axons, which is claimed in the C-P theory. In this study, we provide a theoretical and experimental exploration of the excitability differences between the myelinated motoneurons and unmyelinated muscle fibers. Circuit-Probability simulation concept with distributed-parameter circuit In the previous C-P theory, the target motoneuron is modeled as a parallel RLC circuit. The extracellular medium is simplified as parallel impedance and integrated within the leakage resistor of the parallel RLC circuit. In this paper, the entire muscle tissue, including the target motoneurons and extracellular tissue, is modeled as a distributed-parameter circuit network (Fig. 1). The detailed circuit configuration of each block in this network is determined by its biological features. The myelinated motoneurons are modeled as parallel RLC components (green block). The unmyelinated muscle fibers are modeled as RC components (blue block). The extracellular medium is represented with pure resistive components (𝑹𝑹𝒔𝒔) connecting each block. This circuit network can be further expanded by adding more functional components (e.g. fat tissue and skin as the peripheral circuits) and revised when studying different experimental scenarios (e.g. different configurations and implantation positions of the stimulating electrodes). In this study, a home- made double-side multiple-channel polyimide electrode was implanted in the muscle belly, transversal to the muscle fibers of the Tibialis Anterior (TA) muscle in all experiments, and the force generated by stimulation was measured by a force gauge tied to the ankle (Fig. 2a). Thus, the corresponding circuit network is revised as Fig. 2b. Each electrode site (e1 to e6 refer to the frontside; e1' and e6' refer to the backside) is connected to a node (E1 and E6 refer to the frontside; E1' and E6' refer to the backside). The non-conducting polyimide layer is modeled as broken connections between the frontside and backside. By connecting a current source to an arbitrary electrode sites and a voltage meter to the capacitor in an arbitrary block, the effective voltage waveform at any position in the tissue can be calculated. This voltage waveform will then be used for a probability calculation with a set of parameters, 𝜶𝜶, 𝜷𝜷, and 𝑽𝑽𝑻𝑻𝑻𝑻𝑻𝑻𝑻𝑻𝒔𝒔𝑻𝑻𝑻𝑻𝑻𝑻𝑻𝑻 (parameters to calculate the probability of initiating an action potential), which is the same as the C-P theory. Results Influence of motoneuron-electrode position on voltage waveforms. Apparently, the exact voltage waveform on each block in Fig. 2b is determined by the block position and the electrode sites selected to deliver current. This motoneuron-electrode position will affect the amplitude, shape and the polarity of the voltage waveform on each block. Since there are many motoneuron-electrode combinations, two simplified situations are simulated (Fig. 3a and Fig. 3b), to qualitatively investigate the effect of this motoneuron-electrode position upon the voltage waveform. In Fig. 3a, fixed electrode sites (positive electrode site between P3 and P4; negative electrode site between P9 and P10) are selected to deliver negative-first biphasic square current input. Voltage In another situation (Fig. 3b), two blocks are fixed (P1 and P2) while the spacing between the electrode sites is increased (positive electrode site between P1 and P2 is fixed; negative electrode site changes from E1 to E8). The voltage waveforms on P1 and P2 (Fig. 3d and Fig. 3d') show an opposite changing trend. The peak amplitude of the block outside the electrode sites (P1) and between the electrode sites (P2) increases and decreases with the increasing electrode spacing, respectively. The changing trends of the peak amplitude and polarity (green lines in Fig. 3c, Fig. 3d and Fig. 3d') agree with the results using conventional method of electric field distribution simulation. However, our model can also reveal information in the time domain: voltage waveform, which is critical in determining neuron activation under the C-P theory, but unavailable in previous methods. This voltage waveform can help us understand more phenomena, which will be discussed in the next session. Influence of voltage waveform on force mapping curve: An explanation of the multi-curvature phenomenon. For muscle stimulation using square wave current, it is widely observed that the force mapping curves measured with increasing current amplitude may show multiple curvatures[20][21]. A conventional explanation for this phenomenon is that multiple groups of motoneurons are sequentially recruited when current amplitude is increased. Alternatively, this phenomenon can be quantitatively derived from the C-P theory. Using the same circuit configuration as Fig. 3a, the voltage waveforms of the voltage oscillation induced by RLC components, multiple effective voltage areas (indicated as A and block P1 to P12 in Fig. 3c are shown in Fig. 4a and Fig. 4b, along with an estimated 𝑽𝑽𝑻𝑻𝑻𝑻𝑻𝑻𝑻𝑻𝒔𝒔𝑻𝑻𝑻𝑻𝑻𝑻𝑻𝑻. Due to B in Fig. 4a) will sequentially exceed 𝑽𝑽𝑻𝑻𝑻𝑻𝑻𝑻𝑻𝑻𝒔𝒔𝑻𝑻𝑻𝑻𝑻𝑻𝑻𝑻. when current amplitude is increased, resulting in multiple curvatures in the probability mapping curves (Fig. 4a'). But this phenomenon will not always happen. For the voltage waveforms with only one area exceeding voltage threshold (indicated as C in Fig. 4b), their corresponding probability mapping curves will show a single curvature (Fig. 4b'). These two types of force mapping curves, with and without multiple curvatures, were observed in experiments. waveforms upon different blocks (P1 to P12) are modeled and shown in Fig. 3c. For these different blocks, the voltage amplitude changes, and the polarity also switches. The blocks which are next to the electrode sites (P3 and P4; P9 and P10) have the largest voltage amplitude. Meanwhile, the voltage polarity gradually changes from negative-first to positive-first when the position changes from P1 to P12. Two conclusions can be summarized from the above modeling results and validated by force mapping results in two in vivo experiments (Fig. 4c, Fig. 4c'). First, due to the unknown motoneuron- electrode position after electrode implantation, the stimulating efficiency using the same electrode pair will vary in different experimental trials. In Fig. 4c, the electrode sites with the largest spacing (e1and e6) show minimum stimulation efficiency. However, in another experiment (Fig. 4c'), electrode sites of e1 and e6 show medium stimulation efficiency. Second, the multi-curvature phenomenon will not always happen in the force mapping curves. Whether the curve will have multiple curvatures is determined by the shape of the voltage waveform. The blue curve in Fig. 4c and purple curve in Fig. 4c' clearly show multi-curvature pattern, while other curves don't show such pattern. Activation of multiple groups of motoneurons. As explained above, in contrast to the conventional explanation of multi-curvature force mapping curve as a sequential activation of multiple groups of motoneurons, the activation of a single group of motoneurons can already account for this phenomenon in our theory. We find that the unstable force output during transition periods of additional motorneuron group activation better characterize the recruitment of motornueron groups due to attenuation of voltage over a distance. Two measured force mapping curves (e2e5, orange and e3e4, purple) are selected from Fig. 4c' and plotted in Fig. 5c (blue and orange). There is a sudden force increment accompanied with an abnormally large error bar for the blue curve. We believe that this error bar is a sign of an additional group of motoneurons starting to be recruited, due to the differing voltage acting on spaced out groups of motorneurons. In our experiments (Fig. 1a), the force measured is the summed-up force generated by multiple groups of activated motoneurons. When an additional group of motoneurons starts to be recruited in the stimulation, while the total force will be higher than activating one group alone, the force generated by this second group of motoneurons is still low at this stage. Furthermore, motorneurons that require higher stimulation current to be recruited are necessarily further away and the stimulation strength (voltage upon the motoneurons) will be low, resulting in unstable activation and force generated. Hence, the measured force curve will show a large error bar at this current. Then, when the current further increases, the stimulation strength of this second group of motoneurons will be higher to generate a stable force output, making the error bar recover to a normal level. This explains why this abnormal high error bar only occurs within a certain current range. Fig. 5d shows the force profile measured at this transition range (1000 µA to 1200 µA) of the blue curve in Fig. 5c. When current is high enough (1200 µA), the force profile will recover to a stable condition. Force mapping curves measured in two in-vivo experiments are consistent with the simulation results (Fig. 6c, Fig. 6c'). One force mapping curve was measured at small pulse width range (100-300 µs), and the other is measured at large pulse width range (100-800 µs). Two electrode sites (e1 and e6) of the furthest distance on our polyimide electrode were used. The force mapping curves are the same when e1 delivers positive current or e6 delivers negative current (blue and orange curves, with Fig. 6b' type activation). The recruitment curves are also the same when e1 delivers negative current or e6 delivers positive current (purple and green curves, with Fig. 6b type activation). In each measurement, the four recruitment curves form two groups of different changing trend with increasing pulse width. The flipping of voltage waveform polarity (Fig. 6b and Fig. 6b') causes the voltage waveforms to exceed the threshold in different patterns, so that the two groups have totally different changing trend with increasing pulse width. In addition to the measured force mapping curves, a good curve fitting with distributed-parameter circuit modeling can help us to understand the spatial distribution of these motoneurons within the muscle. Here a simple modeling case is demonstrated. The corresponding electrode pairs of these two curves are modeled in the circuit network (e3e4 in Fig.2b as E1+E1- in Fig. 5a; e2e5 in Fig. 2b as E2+E2- in Fig. 5a). The locations of the two target motoneurons (P1 and P2) are captured to fit the force mapping curves. The probability curves (the simulated curves of activation probability under the C-P theory with respect to input current amplitude or pulse width) are shown in Fig. 5b and Fig. 5b'. At low stimulation current, both E1 and E2 electrode pairs can only stimulate P1. However, when current increases to around 1600 µA, E2 starts to stimulate another motoneuron P2, while E1 has no influence on P2. From this modeling, the relative locations of the two motoneurons with respect to the electrode sites can be roughly estimated. This simple case demonstration shows the potential to apply this model for an in-depth investigation of the biological structure in the future. Influence of input current waveform polarity on motoneuron activation. As shown in Fig. 3c, the polarity of the voltage waveforms at different positions will not be the same. This polarity is not only determined by its position and the polarity of the applied current, but also determined by which electrode is connected to the positive terminal of the current source. It means if the polarity of the input current and the connection of the electrode sites to current source are both reversed, the voltage waveform measured on the specific block will remain the same. One case by modeling is shown in Fig.6. Two electrode sites (E1 and E2) are connected to the current source, which can deliver monophasic square wave current pulses of different pulse widths. The voltage waveform by applying positive pulses from E1 and negative pulses from E2 are the same (Fig. 6b). Similarly, the voltage waveform by applying negative pulses from E1 and positive pulses from E2 are the same. Although the voltage waveforms in Fig. 6b and Fig. 6b' are just of the opposite polarity, the changing trend of the effective voltage area by increasing the pulse width is completely different. Thus, for stimulations using two-electrode configurations, it is important to point out the polarity of current input together with which electrode site is connected to the positive terminal, as even when the same current input is used, results will be different when positive electrode is changed. This issue may not be so prominent in muscle stimulation, because the difference of using different positive electrode only shows in the stimulation efficiency. However, when it comes to the sciatic nerve stimulation using cuff electrode, different positive electrode may stimulate nerve fibers innervating different muscles with the same current input. This will lead to totally different results of activating different muscle groups. Thus, to get consistent results in the experiments, it is important to pay attention to both the polarity of current input and the chosen positive electrode. Independent activation of motoneurons using two-side polyimide electrode. In Fig. 7a, due to the non-conducting polyimide layer, there will be broken connections between the frontside and backside of the electrode. Thus, when current is applied from the frontside electrode sites, motoneurons on the backside won't be activated. In other words, the electrode sites on one side can only cause extremely limited motoneuron activations on the other side of the electrode. A modeling demonstration is shown in Fig. 7a-c. Two electrode sites (E+ and E-) on the frontside of the electrode are selected for current delivery. F1, F2, F3 are three motoneuron positions on the frontside, while B1, B2, B3 are three motoneuron positions on the backside of the electrode. The overall voltage amplitude on the frontside motoneurons is much larger than the backside motoneurons (Fig. 7b, Fig. 7b'). The probability curves show that only with much larger current, the backside motoneurons can be activated by the frontside electrode sites (Fig. 7c). In-vivo experiments proved the independent activation of motoneurons using two-side polyimide electrode. At the beginning of the experiment, force mapping curves for both frontside and backside electrode sites were measured (blue curves in Fig. 7d and Fig. 7d'). Then, the frontside electrode sites were used for a long-time electrical stimulation to induce fatigue. The stimulation lasted for 2.5 min, and the force profile can be found in Fig. 7e. This force profile showed signs of muscle fatigue at the end of the 2.5 min stimulation, as the force dropped to half of the initial value. Right after this 2.5 min stimulation, the force mapping curve using frontside electrode sites was measured (orange curve in Fig. 7d). At that moment, because of the muscle fatigue on the frontside, the force mapping curve was much lower than the one measured in the beginning. Then, the force mapping curve using backside electrode sites was also measured (orange curve in Fig. 7d'). It was only slightly lower than the one measured in the beginning. Lastly, after 5 min resting, the force mapping curve using the frontside electrode sites was measured again (purple curve in Fig. 7d), which showed recovery of muscle fatigue but was still much lower than the measurement in the beginning. Thus, frontside electrode sites can stimulate frontside motoneurons, but can barely stimulate backside motoneurons. It proves our prediction of independent activation of motoneurons using two-side polyimide electrode sites. A theoretical explanation of the difference in excitability of motoneurons and muscle fibers. There are two types of excitable cells in the muscle tissue: motoneurons and muscle fibers. They show distinctive excitability properties in electrical stimulation. In our model, these differences can be theoretically explained and proved by in vivo experiments. The experiments targeting motoneurons and muscle fibers are conducted on healthy muscle and denervated muscle, respectively. Here, two distributed circuits are built to model the healthy muscle and denervated muscle. Since the myelin sheath is proposed as an inductor, the equivalent circuit of a myelinated motoneuron is modeled as a RLC component (Fig. 8a), while an unmyelinated muscle fiber is modeled as a RC component (Fig. 8b). To make a fair comparison, the block representing the target motoneuron or muscle fiber is placed at the same position in the circuit network. The modeling parameters are set based on two reasonable considerations as follows: 1. Due to the lack of myelin sheath, muscle fibers have a larger exposed cell membrane surface and more leakage channels, which can be modeled as a larger capacitor (𝑪𝑪𝟐𝟐>>𝑪𝑪𝟏𝟏) and a lower leakage resistor (𝑹𝑹𝑹𝑹𝟐𝟐<<𝑹𝑹𝑹𝑹𝟏𝟏), respectively. among which the most important is the same threshold voltage, 𝑽𝑽𝑇𝑇ℎ𝑟𝑟𝑟𝑟𝑟𝑟ℎ𝑜𝑜𝑜𝑜𝑜𝑜. 2. It has been reported that the conductance and gating properties of the sodium channels in motoneuron and muscle fiber are nearly the same[22], and muscle fiber sodium channels only require a slightly more negative potential to activate than motoneuron sodium channels[23][24]. Therefore, in our modeling, the motoneurons and muscle fibers share the same parameters for probability calculus, Firstly, the voltage waveforms are compared when the same negative monophasic current pulse (100 µs pulse width) is applied (Fig. 8c1). The voltage waveform of the RC component shows a typical charging and discharging curve as a capacitor, while the voltage waveform of the RLC component shows a voltage oscillation, resulting in a much higher peak amplitude. Thus, to exceed the threshold voltage and activate the muscle, the threshold current required for RLC component is much lower than RC component. The probability mapping curves by changing current amplitude in Fig. 8c2 show the difference in threshold current. The probability mapping curves with increasing pulse width in Fig. 8c3 show that RC circuit requires a much higher current to achieve the same force. This phenomenon is also observed in previous report[25] and confirmed by our in vivo experiments (Fig. 8c4). To achieve the similar force output, healthy muscle requires only 1 mA while the denervated muscle requires 10 mA. Then the changing trend of the voltage waveforms with increasing pulse width is compared in Fig. 8d1 and Fig. 8d2. The voltage waveforms of the RC component show a simple charging and discharging process as a capacitor. Although the slope of the curve, which represents the charging rate, will vary with the circuit parameters, the general shape will not change. Thus, the effective voltage area, which is the part exceeding 𝑽𝑽𝑻𝑻𝑻𝑻𝑻𝑻𝑻𝑻𝒔𝒔𝑻𝑻𝑻𝑻𝑻𝑻𝑻𝑻, will always have a monotonically increasing trend. As a result, the corresponding probability mapping curves (Fig. 8d3) and measured force mapping curves (Fig. 8d4) also increase with respect to pulse width monotonically. However, the changing trend of the effective voltage area of the RLC component is quite complex. Many factors, such as current amplitude, circuit parameter and extracellular environment, can induce nonlinear effect upon the effective voltage area and finally result in various probability mapping patterns, which is investigated in details in our previous C-P theory[19].Fig. 8d2 just shows a typical voltage waveform of RLC component. With different current amplitudes, the effective voltage waveform will be completely different. In summary, due to the lack of myelin sheath, muscle fiber is not very easily stimulated being pure RC, but because of the branching motoneurons coordinating the impulse it is able to be activated with lower current in healthy muscles. Meanwhile, the force mapping curves by changing pulse width in muscle fiber stimulation always follow a monotonically increasing trend while the force mapping curves of motoneuron stimulation will have a complex pattern[19]. Discussion A static system-level distributed-parameter circuit to study electrical stimulation on muscle tissue. In this paper, instead of treating the muscle tissue as a black box and trying to get some simplified descriptions about the relationships between input and output, we thoroughly studied the interaction between the stimulation electrodes and muscle tissue. This interaction is highly affected by all parameters in the system, including the motoneuron-electrode position, input current waveforms (amplitude, pulse width, frequency, and polarity), circuit properties of the tissue components (motoneuron, muscle fiber, and extracellular medium), and the ion channel properties (𝜶𝜶, 𝜷𝜷, and 𝑽𝑽𝑻𝑻𝑻𝑻𝑻𝑻𝑻𝑻𝒔𝒔𝑻𝑻𝑻𝑻𝑻𝑻𝑻𝑻). Therefore, a simplified general description of the relationship between input and output does not exist. No matter what kind of simplified conclusions are summarized from the experiment observations, exceptions will always exist. Here, we demonstrate a modified distributed-parameter circuit to study a specific electrode implantation: intramuscular implantation of a double-side multiple-channel polyimide electrode. There are many electrode designs for muscle stimulation, including epimysial electrode and skin surface electrode. These electrodes also come in different structures, like flexible strip electrodes, wire electrodes and large-area patch electrodes. By including circuit components representing additional tissue layers (fat layer, skin) and properly assigning the motoneuron-electrode positions, modified distributed-parameter circuit can be applied to study all these implantations. A potential solution to achieve sustained force output with reduced muscle fatigue in electrical stimulation. Our distributed-parameter circuit can provide guidance to solve practical issues in electrical stimulation. One example is to use double-side electrode stimulation for less muscle fatigue. The goal of Functional Electrical Stimulation (FES) is to achieve meaningful functions using precisely controlled electrical stimulation. However, electrical stimulation causes muscle fatigue much more easily than the voluntary controlled movements. The fast kick-in of muscle fatigue is an obstacle to achieve some meaningful functions, such as holding and gripping. In this paper, we have proved the independent recruitment of motoneurons using double-side polyimide electrode. Even when the frontside motoneurons are severely fatigued with the frontside electrodes delivering electrical stimulation, the backside motoneurons are barely affected and can still generate large force with the backside electrodes. The independent recruitment of motoneurons results in independent fatigability within a muscle. If we expand this concept to a meaningful function, such as a holding movement, the frontside and backside electrodes can be controlled to alternatively deliver the electrical stimulation (Fig. 9). By alternating the side of electrode sites during a long stimulation, the muscle fibers and motoneurons on one side can rest and recover when the other side is recruited for force output. Potential applications of distributed-parameter circuit model to study electrical stimulation on other neural tissue. Apart from muscles, this distributed-parameter circuit can also be applied to other excitable tissues, including the Peripheral Nervous System (PNS) and the Central Nervous System (CNS). In the distributed-parameter circuit description for the PNS and CNS, the muscle fibers will be removed, and much more neuron distribution will be added. Further study using our distributed-parameter circuit will help to elucidate important issues in the PNS and CNS electrical stimulation. An example is to use high frequency localized electrical stimulation for selective activation. Selective stimulation of different nerve fascicles on the sciatic nerve can enable functional activation of various muscle groups. In addition, localized stimulation in the spinal cord can eliminate unwanted disturbance on other functional areas. In the previous research, without this system-level model, researchers tended to use experiment observations to judge the efficiency of electrical stimulation. Now we know the stimulation effect under a specific condition is affected by so many parameters in the system. Since in the way of experiment observations, only a certain range of parameters can be applied, this may give rise to a segmented conclusion about this specific stimulation. This segmented conclusion may neglect the complicated interaction between the stimulation electrodes and excitable tissue. Here, we take the study of tripolar stimulation on the sciatic nerve as an example, to demonstrate how segmented conclusions can arise when only parameters of a certain range are applied during the experiment observations. In Fig. 10, the two outer electrode sites are connected to the positive terminal of the current source, and the middle electrode site is connected to the negative terminal. Positive monophasic input current is delivered using this tripolar structure. With the knowledge about voltage waveform polarity in Fig. 6, we now know that it is the middle electrode site which is delivering negative input current that stimulates the nerve axons to generate action potentials. When these action potentials travel to the electrode site located at the lower stream of the nerve axons, they will be distorted by the external stimulation at that location. Whether these action potentials can successfully pass through the lower-stream electrode site depends on the external stimulation at that location: they can pass through when the amplitude of the external stimulation is smaller than the action potential (Fig. 10a), and they will be blocked when the amplitude of the external stimulation is larger (Fig. 10b). However, it is also possible that when the action potentials arrive at the lower-stream electrode site, the external stimulation of a short pulse width has already finished. Then, in this case, the action potentials will pass through at both small and large amplitude (Fig. 10c, Fig. 10d). Thus, if we use force measurement as a quantification of the action potentials passing through the lower-stream electrode site, the force curve with respect to the current amplitude will show totally different patterns when different pulse width is used. A new explanation of cathodal make, cathodal break, anodal make, and anodal break stimulation phenomena. It has been observed that cardiac tissue can be electrically stimulated with the onset (make) or termination (break) of an input current that is delivered with either a negative (cathodal) or positive (anodal) electrode[26][27]. These phenomena can be easily explained with our model. In the case of unipolar stimulation configuration, where only one electrode delivers input current, the effectiveness of cathodal make can be intuitively understood, as it will depolarize the tissue surrounding the electrode. However, considering the dramatic resonance caused by the RLC component, cathodal break, anodal make, and anodal break may also generate voltage waveforms with a certain area sufficiently negative to exceed the voltage threshold. The gap between static and dynamic distributed-parameter circuit model. Inherited and expanded from the C-P theory, we still assigned static values to parameters of both the circuit components (resistance, inductance, and capacitance) and probability calculation (𝜶𝜶, 𝜷𝜷 and 𝑽𝑽𝑻𝑻𝑻𝑻𝑻𝑻𝑻𝑻𝒔𝒔𝑻𝑻𝑻𝑻𝑻𝑻𝑻𝑻). By using static parameters of circuit components, we indeed assumed that the electrical properties of the muscle tissue remain unchanged during applied electrical stimulation. Similarly, by using static parameters of probability calculation, we assumed that external electrical stimulation won't affect the resting potential and threshold voltage of motoneurons. This adoption of static parameters limits our capability to study some specific issues. One example is the quantitative description of muscle fatigue during electrical stimulation, which is characterized as change in force output. During muscle fatigue, the same current input is applied, but the output force is different. It means something must have changed, either the generated voltage waveform affected by circuit parameters, or the ion channel response to the voltage waveform which is affected by probability calculation parameters. Another limitation caused by the adoption of static parameters is that we cannot quantitatively study how an external electrical stimulation interacts with a generated action potential. With the static circuit parameters, we can only study how the external stimulation initiates an action potential. Unlike the natural action potentials, the action potential initiated by the external electrical stimulation may be distorted by this external stimulation. For the external stimulation that lingers on after the action potential is already initiated, it will interact with the movements of local ions across the neuron membrane and result in a distorted action potential. Thus, to quantitatively study these dynamic processes in electrical stimulation, further studies are required to establish a bridge between the parameters in our model and the dynamic change of muscle tissue. Reason for using voltage waveforms simulated with a distributed-parameter circuit, instead of electric field distribution simulated with the existing Finite Element Modeling (FEM). The cross- membrane electric field opens the ion channels. Considering the neuron membrane is partially permeable to ions, charged ions will accumulate on both sides of the neuron membrane. Thus, the electric field generated by these local ions is perpendicular to the membrane. To model the effect of this perpendicular cross-membrane electric field (𝑬𝑬) on the neuron, we calculate the voltage across the the capacitor can be used to characterize the electric field perpendicular to the membrane, which really opens an ion channel. However, the electric field simulated with the existing FEM is only determined by the physical boundary condition and electrode positions, which is irrelevant to the electric field perpendicular to the neuron membrane. capacitor (𝑽𝑽) in our model. This 𝑽𝑽 is the path integral of the electric field 𝑬𝑬 in the direction perpendicular to the capacitor plates (neuron membrane) as described by the equation ∆𝑽𝑽=𝑬𝑬∗∆𝑻𝑻 (∆𝑽𝑽 is a change of 𝑽𝑽, ∆𝑻𝑻 is a change of membrane thickness). Thus, in our model, the voltage across nonlinear relationship between 𝑽𝑽(𝒕𝒕) (cross-membrane voltage at time point t) and 𝑰𝑰(𝒕𝒕) (input current at time point t). In response to an external input current 𝑰𝑰(𝒕𝒕) , the local ions on both sides of the neuron assuming a linear relationship between 𝑽𝑽(𝒕𝒕) and 𝑰𝑰(𝒕𝒕) . membrane will move. This local ion movement is not only affected by the charging and discharging process of the neuron membrane, but also affected by the surrounding tissue environment. That's why we build a distributed-parameter circuit network to account for the influence of the other parts of the tissue. However, in the existing FEM, the tissue is modeled in a medium. In this way, the existing FEM is In our model, by calculating the voltage across the capacitor (neuron membrane), we can model the Methods Animal statement. All experiments were conducted according to protocols approved by the Institutional Animal Care and Use Committee at the National University of Singapore. Preparation of denervated muscles. Sprague-Dawley rats (around 450 g) were used for the sciatic nerve transaction. Anesthesia was induced with isoflurane. Buprenorphine was injected for pain relief before the surgery. After the rat was anesthetized, a shaver was used to gently remove the fur on the left leg. Then, the skin was disinfected with 70% ethanol wipes, and an incision was made to expose the bicep femoris muscle. An incision was then made on this bicep femoris muscle to expose the sciatic nerve. The sciatic nerve was transacted on the site before it branched into three smaller nerves. Then the bicep femoris muscle and the skin were both sutured back. Right after the surgery, enrofloxacin was injected. During the first five days after the nerve transaction surgery, buprenorphine was injected twice a day and enrofloxacin was injected once a day. Interface implantation. The homemade flexible interface was folded to form a loop on the tip. A suture was threaded through this loop. Then, this suture was threaded through the center of the exposed TA muscle belly. Pulling by the suture, the interface was also threaded into the muscle belly. The interface was sutured to the muscle surface for fixation. Then, the skin was sutured back. Electrical stimulation. A-M SYSTEMS model 4100 isolated high-power stimulator was used for electrical stimulation. Every one second, a train of 10 negative-positive biphasic pulses of 60 Hz was applied. In the experiment of comparing four waveforms, positive-negative biphasic pulses, negative- positive biphasic pulses, positive monophasic pulses, and negative monophasic pulses were applied. Force data collection and analysis. The anesthetized rat was fixed on a stand, and the ankle of the left leg was connected to a dual-range force sensor (Vernier). This force sensor was connected to a laptop through a data acquisition device (National Instruments). LabView (National Instruments) was used for on-site result visualization during the measurements. After the measurements, MATLAB (MathWorks) was used for data analysis. The recorded forces exceeding a certain amplitude threshold were used for assessing the strength of the excitability. If the forces were below this amplitude threshold, then the muscle was considered as not activated. Distributed-parameter circuit simulation. The simulation was performed on MATLAB (MathWorks). Firstly, a circuit description was performed in Simulink (MathWorks). Then, current inputs of different waveforms were recursively fed to the circuit model, and the voltage responses of the targeted RC or RLC component were collected. Lastly, these voltage responses were fed into the probability equation to Acknowledgements calculate the probability of excitation under these current inputs. The circuit parameters and probability calculation parameters are shown in Table. 1 and Table. 2. This work was supported by the following grant from the National Research Foundation: Competitive Research Project 'Peripheral Nerve Prostheses: A Paradigm Shift in Restoring Dexterous Limb Function' (NRF-CRP10-2012-01), and the grant from the HIFES Seed Funding: 'Hybrid Integration of Flexible Power Source and Pressure Sensors' (R-263-501-012-133). We would like to thank for the experiment setup support from Han Wu, Li Jing Ong and Wendy Yen Xian Peh. We also would like to thank for the animal experiment support from Gammad Gil Gerald Lasam. [1] Anderson, K. D. Targeting recovery: priorities of the spinal cord-injured population. J. Neurotrauma. 21(10), 1371-1383 (2004). [2] Galvani, L. De viribus electricitatis in motu musculari commentarius. De Bononiensi Scientiarum et Artium Instituto atque Academia Commentarii, 7, 363 -- 418 (1791). [3] Hodgkin, A. L., and Andrew F. H. A quantitative description of membrane current and its application to conduction and excitation in nerve. J. Physiol. 117, no. 4: 500-544 (1952). [4] McNeal, D. R. Analysis of a model for excitation of myelinated nerve. IEEE Trans Biomed Eng. 4, 329- 337 (1976). [5] Raspopovic, S., Capogrosso, M., & Micera, S. A computational model for the stimulation of rat sciatic nerve using a transverse intrafascicular multichannel electrode. IEEE Trans Neural Syst Rehabil Eng 19(4), 333-344 (2011). [6] Raspopovic, S., Petrini, F. M., Zelechowski, M., & Valle, G. Framework for the development of neuroprostheses: from basic understanding by sciatic and median nerves models to bionic legs and hands. Proc. IEEE. 105(1), 34-49 (2017). [7] Pelot, N. A., Behrend, C. E., & Grill, W. M. Modeling the response of small myelinated axons in a compound nerve to kilohertz frequency signals. J Neural Eng. 14(4), 046022 (2017). [8] Capogrosso, M. et al. A computational model for epidural electrical stimulation of spinal sensorimotor circuits. J. Neurosci. 33(49), 19326-19340 (2013). [9] McIntyre, C. C., & Grill, W. M. Extracellular stimulation of central neurons: influence of stimulus waveform and frequency on neuronal output. J. Neurophysiol. 88(4), 1592-1604 (2002). [10] Dehghani, N., Bédard, C., Cash, S. S., Halgren, E., & Destexhe, A. Comparative power spectral analysis of simultaneous elecroencephalographic and magnetoencephalographic recordings in humans suggests non-resistive extracellular media. J Comput Neurosci. 29(3), 405-421 (2010). [11] Bédard, C., Rodrigues, S., Roy, N., Contreras, D., & Destexhe, A. Evidence for frequency-dependent extracellular impedance from the transfer function between extracellular and intracellular potentials. J Comput Neurosci. 29(3), 389-403 (2010). [12] Reimann, M. W. et al. A biophysically detailed model of neocortical local field potentials predicts the critical role of active membrane currents. Neuron, 79(2), 375-390 (2013). [13] Gomes, J. M. et al. Intracellular impedance measurements reveal non-ohmic properties of the extracellular medium around neurons. Biophys. J. 110(1), 234-246 (2016). [14] Aldrich, R. W., & Stevens, C. F. Voltage-dependent gating of single sodium channels from mammalian neuroblastoma cells. J. Neurosci. 7(2), 418-431 (1987). [15] Zagotta, W. N., & Aldrich, R. W. Voltage-dependent gating of Shaker A-type potassium channels in Drosophila muscle. J. Gen. Physiol. 95(1), 29-60 (1990). [16] Zagotta, W. N., Hoshi, T., & Aldrich, R. W. Gating of single Shaker potassium channels in Drosophila muscle and in Xenopus oocytes injected with Shaker mRNA. PNAS, 86(18), 7243-7247 (1989). [17] White, J. A., Rubinstein, J. T., & Kay, A. R. Channel noise in neurons. Trends Neurosci. 23(3), 131-137 (2000). [18] Mino, H., Rubinstein, J. T., & White, J. A. Comparison of algorithms for the simulation of action potentials with stochastic sodium channels. Ann Biomed Eng, 30(4), 578-587 (2002). [19] Hao, W. et al. Unveiling Stimulation Secrets of Electrical Excitation of Neural Tissue Using a Circuit Probability Theory. Submitted to Nature. [20] Crago, P. E., Peckham, P. H., & Thrope, G. B. Modulation of muscle force by recruitment during intramuscular stimulation. IEEE Trans Biomed Eng. 12, 679-684 (1980). [21] Popovic, D., Baker, L. L., & Loeb, G. E. Recruitment and comfort of BION implanted electrical stimulation: implications for FES applications. IEEE Trans Neural Syst Rehabil Eng. 15(4), 577-586 (2007). [22] Campbell, D. T., & Hille, B. E. R. T. I. L. Kinetic and pharmacological properties of the sodium channel of frog skeletal muscle. J. Gen. Physiol., 67(3), 309-323 (1976). [23] Pappone, P. A. Voltage‐clamp experiments in normal and denervated mammalian skeletal muscle fibres. J. Physiol. 306(1), 377-410 (1980). [24] Weiss, R. E., & Horn, R. Functional differences between two classes of sodium channels in developing rat skeletal muscle. Science, 233(4761), 361-364 (1986). [25] Peckham, P. H., & Knutson, J. S. Functional electrical stimulation applications. Annu. Rev. Biomed. Eng., 7, 327-360 (2005). [26] Dekker, E., Direct current make and break thresholds for pacemaker electrodes on the canine ventricle. Circ. Res. 27(5), pp.811-823 (1970). [27] Wikswo Jr, J. P., Lin, S. F., & Abbas, R. A. Virtual electrodes in cardiac tissue: a common mechanism for anodal and cathodal stimulation. Biophys. J. 69(6), 2195-2210 (1995). for neuromuscular Figure 1 Distributed-parameter circuit simulation concept. Rs represents resistor of extracellular media. RLC components (green blocks) represent motoneurons distributed in the healthy innervated muscle tissue. RC components (blue blocks) represent muscle fibers. Figure 2 In vivo measurement setup and the corresponding distributed-parameter circuit concept. (a) Our home-made double-side multiple-channel polyimide electrode is sutured in the muscle belly, transversal to muscle fibers of the Tibialis Anterior (TA) muscle. When electrical stimulation is delivered to the TA muscle, the leg can freely kick forward. Force is measured from the ankle. (b) The corresponding distributed-parameter circuit with broken connections between two sides of electrode sites representing the insulating polyimide substrate, and RLC components (green blocks) representing distributed motoneurons. Figure 3 Influence of motoneuron-electrode position on voltage waveform. (a) Distributed-parameter circuit of fixed electrode site location and changing targeted motoneuron positions (green blocks P1- P12). (b) Distributed-parameter circuit of fixed positive electrode site location and changing negative electrode site locations (E1-E8). Two targeted motoneuron positions (P1, P2) are studied, P2 is between the electrode sites, and P1 is outside. (c) Voltage waveforms on motoneuron positions in (a). (d, d') Voltage waveforms on P1 (d) and P2 (d') with negative electrode positions in (b). Figure 4 Force mapping curves with multiple curvatures can be explained with the voltage waveforms. (a, a') Voltage waveforms (a) and corresponding probability curves (a') of P1-P6 in Fig. 3c. (b, b') Voltage waveforms (b) and corresponding probability curves (b') of P7-P12 in Fig. 3c. For voltage waveforms with two areas exceeding the threshold (area A and area B), the probability curves have two corresponding curvatures. For voltage waveforms with only one area exceeding the threshold (area C), the probability curves also only show one curvature. (c, c') Quantification of force at different current measured in two in-vivo experiments. Data are means ± s.d. (n=15 per group). Figure 5 Activation of multiple groups of target motoneurons. (a) Distributed-parameter circuit with two groups of electrode sites (E1+ and E1-, E2+ and E2-) and two targeted motoneuron positions (P1, P2). (b, b') Simulation results. Probability curves of two groups of electrode sites on P1 and P2 separately (b). Probability curves of the summed stimulation effects on P1 and P2 using two groups of electrode sites (b'). (c) Quantification of force at different current. The two curves using different groups of electrodes have a crossing point. (d) Force profile at four current corresponding to the transition points in (c). Data are means ± s.d. (n=15 per group). Figure 6 Influence of input current waveform polarity on motoneuron activation. (a) Distributed- parameter circuit with two electrode sites (E1 and E2) and a targeted motoneuron (green block). (b, b') Simulated voltage waveforms of different pulse width. E1 delivers positive monophasic current or E2 delivers negative monophasic current (b). E1 delivers negative monophasic current or E2 delivers positive monophasic current (b'). (c) Quantification of force at different pulse width in two experiments. Data are means ± s.d. (n=15 per group). Figure 7 Independent activation of motoneurons using double-side polyimide electrode. (a) Distributed-parameter circuit with broken connections to represent the non-conducting polyimide substrate. E- and E+ are electrode sites on the frontside of the polyimide electrode. F1, F2, F3 are motoneurons on the frontside, and B1, B2, B3 are motoneurons on the backside of the polyimide electrode. (b, b') Voltage waveforms on F1-F3 (b) and B1-B3 (b') when current input is delivered from frontside electrode sites. The scale of the front motoneuron waveforms are much larger than the back motoneuron waveforms. (c) The corresponding probability curves. (d, d') In-vivo measurement results. Quantification of force at different current during frontside stimulation (d) and backside stimulation (d'). (e) Force profile of 2.5 min electrical stimulation to induce fatigue on frontside. Data are means ± s.d. (n=15 per group). Figure 8 The inductive myelin sheath on motoneurons accounts for the huge difference in excitability of motoneurons and muscle fibers. (a) Distributed-parameter circuit with RLC component (green block) representing the target motoneuron in healthy muscle stimulation. (b) Distributed- parameter circuit with RC component (blue block) representing the target muscle fiber in denervated muscle stimulation. (c1, c2) Simulated voltage waveforms (c1) and probability curves (c2) on RLC component and RC component (c3, c4) Simulated probability curves (c3) and quantification of force (c4) at small pulse width. (d1, d2) Simulated voltage waveforms of RC (d1) and RLC (d2) component at different pulse width (d3, d4) Simulated probability curve (d3) and quantification of force of a denervated muscle (d4) at large pulse width. Data are means ± s.d. (n=15 per group). Figure 9 Illustration of a long-duration current waveform using double-side electrode stimulation for reduced muscle fatigue. The frontside and backside electrode sites are controlled to alternatively deliver the stimulation current. Figure 10 Illustration of tripolar stimulation on the sciatic nerve. The two outer electrode sites are connected to the positive terminal of the current source, and the middle electrode site is connected to the negative terminal. A positive monophasic input current is applied using this tripolar structure. (a, b) Large pulse width. Action potentials generated by the negative electrode site can pass through the lower-stream positive electrode when the current amplitude is small (a). Action potentials cannot pass through when the current amplitude is large (b). (c, d) Small pulse width. Action potentials can pass through when the current amplitude is small (c) and large (d). Figure 11 Concept of conventional computational model and our model to study excitability of tissue. Table 1 Parameter table of components used in the distributed-parameter circuit. Table 2 Parameter table used in distributed-parameter circuit simulation and probability calculation. Extended Figure 1 Influence of input biphasic current polarity on motoneuron activation. (a) Distributed-parameter circuit with two electrode sites (E1 and E2) and a targeted motoneuron (green block). (b) Simulated voltage waveforms of different pulse width. E1 delivering positive-negative biphasic current and E2 delivering negative-positive biphasic current share the same voltage waveform. E1 delivering negative-positive biphasic current and E2 delivering positive-negative biphasic current share the same voltage waveform. (c) Quantification of force at different pulse width in two experiments. According to the simulation in (b), the four measured recruitment curves should come in two groups (e1 positive-negative and e6 negative-positive in one group, e1 negative-positive and e6 positive-negative in another group). However, the two measurements show a single group. (d) To investigate the reason for showing a single group in recruitment curves, instead of two, simulation was carried out to see how the two groups of possibility curves change with stimulation current. The probability curve shows that the voltage waveform in response to positive-negative, and negative- positive current input have a slightly different changing trend with increasing current, and there is a crossing point. (e) Recruitment curves of e1 delivering positive-negative and negative-positive current. The two recruitment curves overlap in a large range of current. This explains why only a single group of recruitment curves is observed in (c). Extended Figure 2 Shift of force mapping curves due to muscle fatigue. Quantification of force at different current in the first round of recruitment curve measurement (first panel). Then, the recruitment force using each electrode site combination was measured again. For each electrode site combination, the recruitment curve shifted towards larger current and smaller force output, due to accumulated muscle fatigue. Data are means ± s.d. (n=15 per group). Extended Figure 3 Shift of recruitment curves due to muscle fatigue. In both testing, three rounds of recruitment measurement were carried out. In each round, the current was changed from large to small, and then small back to large. In both testing, the recruitment curves shift towards smaller force output, due to accumulated muscle fatigue. Extended Figure 4 Recruitment curves of denervated muscles using 10 mA current. The recruitment curves were measured on denervated muscle, using both sides of the polyimide electrode. Extended Figure 5 Electric potential distribution generated by two point charges. Electric potential is calculated for positions in the line of the two charges. The amplitude of the generated electric potential is the largest near the two charges. The polarity of the electric potential changes at the center of the two charges. This distribution can help to understand the change of voltage waveforms generated at different location in our model.
1501.05174
2
1501
2015-04-19T14:13:12
Characterizing the neural correlates of reasoning
[ "q-bio.NC" ]
The brain did not develop a dedicated device for reasoning. This fact bears dramatic consequences. While for perceptuo-motor functions neural activity is shaped by the input's statistical properties, and processing is carried out at high speed in hardwired spatially segregated modules, in reasoning, neural activity is driven by internal dynamics, and processing times, stages, and functional brain geometry are largely unconstrained a priori. Here, it is shown that the complex properties of spontaneous activity, which can be ignored in a short-lived event-related world, become prominent at the long time scales of certain forms of reasoning which stretch over sufficiently long periods of time. It is argued that the neural correlates of reasoning should in fact be defined in terms of non-trivial generic properties of spontaneous brain activity, and that this implies resorting to concepts, analytical tools, and ways of designing experiments that are as yet non-standard in cognitive neuroscience. The implications in terms of models of brain activity, shape of the neural correlates, methods of data analysis, observability of the phenomenon and experimental designs are discussed.
q-bio.NC
q-bio
Characterizing the neural correlates of reasoning David Papo1,2* 1 Laboratory of Biological Networks, Center for Biomedical Technology, Universidad Politécnica de Madrid 2 LPC Institute, Madrid *Correspondence: David Papo, Laboratory of Biological Networks, Center for Biomedical Technology, Universidad Politécnica de Madrid, Calle Ramiro de Maeztu, 7, 28040 Madrid, Spain e-mail: [email protected] www: https://davidpapo.wordpress.com/ The brain did not develop a dedicated device for reasoning. This fact bears dramatic consequences. While for perceptuo-motor functions neural activity is shaped by the input’s statistical properties, and processing is carried out at high speed in hardwired spatially segregated modules, in reasoning, neural activity is driven by internal dynamics, and processing times, stages, and functional brain geometry are largely unconstrained a priori. Here, it is shown that the complex properties of spontaneous activity, which can be ignored in a short-lived event-related world, become prominent at the long time scales of certain forms of reasoning which stretch over sufficiently long periods of time. It is argued that the neural correlates of reasoning should in fact be defined in terms of non-trivial generic properties of spontaneous brain activity, and that this implies resorting to concepts, analytical tools, and ways of designing experiments that are as yet non-standard in cognitive neuroscience. The implications in terms of models of brain activity, shape of the neural correlates, methods of data analysis, observability of the phenomenon and experimental designs are discussed. Keywords: cognitive neuroscience, reasoning, scaling, non-stationarity, non-ergodicity, characteristic scales, observation time, resting brain activity. INTRODUCTION Consider an individual trying to solve a problem, and reasoning for ten minutes before attaining a solution. Take the middle five minutes. Clearly, though containing no behaviourally salient event, these five minutes represent a genuine, indeed rather general, instance of reasoning. What do we know about the brain regime far from its conclusion? Can we use this regime to predict a solution, and a solution to retrodict this regime? Here, I concentrate on a form of reasoning, of which the above scenario constitutes an example, which can broadly be defined as "thinking in which there is a conscious intent to reach a conclusion and in which methods are used that are logically justified" [1], with no a priori assumption on the type of reasoning process that may take place during it. It is argued that finding the generic properties of this form of reasoning entails addressing the following fundamental issues: What are reasoning's temporal and spatial scales? When is a given observation time sufficient? How should we integrate the information contained in various reasoning episodes? A MINI LITERATURE REVIEW The neural correlates of reasoning have traditionally been expressed in terms of brain spatial coordinates. Early neuropsychological work viewed reasoning as emerging from global brain processing [2], consistent with evidence indicating that it is negatively affected by diffuse brain damage [3]. Neuroimaging studies have framed the neural correlates of reasoning local functionally specialized brain activity, either by taking a normative or by fractionating it into sub-component processes [11-14]. The results often lack specificity to reasoning [15]. Most terms of reasoning approach [4-10], in to importantly though, these investigations provide a static characterization of reasoning. The neuroimaging literature mostly focused on short- term and normative forms of reasoning [9,16-18]. This minimizes variability in reasoning episode length and allows segmenting reasoning episodes into separable chunks, but does that at the price of limitations in the phenomenology and ecologic value of its stimuli. Some neuroimaging [19,20] and electrophysiological [21- 29] studies examined more ecological forms of reasoning, viz. even electrophysiological studies, despite optimal temporal resolution, adopted an event-related perspective, concentrating on activity occurring few seconds before insight emergence , which only documents the outcome of the reasoning process, not the process itself. However, problems insight [30]. Event-related neural activity associated with the solution of riddles with insight was found to be related to properties of preceding resting activity [26,27]. These studies had the remarkable merit of using spontaneous brain activity to characterize reasoning, but in essence provided a comparative statics description. Although some behavioural studies as a dynamical process [31], a comparable neurophysiological characterization incomplete. Altogether, the research accomplished so far has generally not looked at reasoning as a dynamical process, and produced either time-averaged frames or discrete maps in partial time cuts. treated reasoning is still THE PROBLEM(S) WITH REASONING The generalized form of reasoning considered in this study comes in episodes offering scant behaviourally salient events with no characteristic temporal length. Each episode is a non-reproducible instance, as reasoning task can be carry out in multiple ways. Brain activity l correlates of reasoning Papo Characterizing the neura Papo Characterizing the neural correlates of reasoning l correlates of reasoning Papo Characterizing the neura Papo Characterizing the neura l correlates of reasoning associated with reasoning is not event-related, and many neurophysiological processes interact in a wide range of spatial and temporal scales. These phenomena can all be traced back to a basic fact: the brain did not develop a dedicated device for reasoning. Hardwired partially segregated modules ensure that perceptuo-motor functions are carried out at great speed, with stereotyped duration and time-varying profile, and identifiable stages, largely determined by input statistical properties. Reasoning, on the contrary, is an internally-driven dynamics: processing times, stages, and functional brain geometry are largely unconstrained. Considering these extraordinary challenges, can we still find general reasoning properties, over and above specific task demands and individual differences? What sort of process is reasoning in its general form? Is it a series of simpler reasoning cycles? Can we segment it into stages? What are the best neural variables and tools to make these properties observable? CHARACTERIZING THE REASONING PROCESS tools accordingly. For Robust characterizations of reasoning should incorporate stylized facts, i.e. properties consistently appearing on different subjects and in different periods of time, and select analytical instance, perceptual response sensitivity to incoming signals, stability against noise, and minimal dependence on initial conditions favour tools capturing transient dynamics, which naturally reproduce these properties under appropriate conditions, over tools handling asymptotic activity, which fail to do so [32]. from instability and and protection Reasoning's relative inefficiency that optimal circuitry may need constant suggest reconstruction interference, summoning protracted support of energetically costly long-range communications. Thus, reasoning may be a sort of resonant regime, where functional efficiency would be achieved with specific, though unstable, spatio- temporal patterns, and should be studied with tools capable of spatially-extended dynamic transients. extracting REASONING DYNAMICS Each cognitive process can be translated in dynamical terms and corresponding aspects of neural activity. are processes Perceptual quasi- stereotyped short duration processes. The brain can prima facie be modelled as an excitable medium: perturbations above a threshold induce a dynamical cycle, before the system reverts to its initial silent state. relaxational, the brain Learning too is a relaxational process: following a gradient dynamics, the environment's statistical relationships, by representing them in terms of its functional connectivity. Cycles can be of much longer duration and non-trivial shape than perceptual ones. The dynamics is dominated by fluctuations much shorter than the whole process. incorporates Reasoning may not be purely relaxational. The corresponding neural activity is a fluctuation-dominated endogenously modulated spontaneous brain activity, with no clear gradient, and no single instant summarizing the entire process. The reasoning scientist's world is considerably more complex than the event-related short time scale one of perception. It is through the generic properties of spontaneous activity's long time scales that robust reasoning properties should be formulated [33]. THE STARTING POINT: SPONTANEOUS BRAIN ACTIVITY Spontaneous activity can be thought of as a data bank of cortical states, continuously reedited across the cortex [34]. This re-editing process contains rich non-random spatio-temporal structure [35-40]. The building blocks of this structure are fluctuations which the brain, as all dissipative out-of-equilibrium systems, generates even for fixed control parameter values and in the absence of external stimuli, and which constitute the trademark of its functioning. The dynamics is intermittent, with alternating laminar and turbulent phases [41,42], weakly non-ergodic, i.e. some phase space regions take extremely long times to be visited, and shows aging, i.e. temporal correlations depend on the observation time [43]. Various aspects of spontaneous activity display similar properties at all temporal and spatial scales [44-51]. Self-similarity break-down [52,53], with Gaussian low-frequency and non-Gaussian high- frequency fluctuations [54] were also reported. Understanding fluctuations We can imagine brain activity as the motion of a random walker, making steps of a length taken from some distribution, at times taken from some other distribution or, equivalently, of a macroscopic particle diffusing in a liquid, subject to viscous friction and to an additive random force [55]. from a Gaussian distribution The relationship between these two forces' time scales determines how microscopic fluctuations produce observable macroscopic properties. In the equilibrium world of perceptual scientists, the brain makes steps taken and whose correlations, much faster than the friction time scale, have no macroscopical effect. Reasoning scientists live far from equilibrium: random fluctuations are no faster than the friction term and show long-lasting correlations, which renormalize, becoming macroscopically detectable [56]. Complex fluctuations reveal the particle's navigation 'style', e.g. how travelled distances and the times to reach a given target scale with time [57]. They also allow deducing a system's characteristic temporal scales. For an equilibrium temporal autocorrelation decay, the correlation length, i.e. the value ߦ that makes the autocorrelation ܥሺݐ = ߦሻ = 0, or the correlation time ߬஼ = ׬ ܥሺݐሻ݀ݐ , endow the process with a unique temporal scale. However, at long time scales, the presence of scaling indicates that the brain relaxes more system with exponential క ଴ l correlates of reasoning Papo Characterizing the neura Papo Characterizing the neural correlates of reasoning l correlates of reasoning Papo Characterizing the neura Papo Characterizing the neura l correlates of reasoning slowly than an exponential and fluctuates at all scales [58- 60]. Both ߦand ߬஼ may diverge, and a characteristic time ceases to exist. Temporal scales are characterized by some relationship between them, which can be treated as a dynamical system, relating the behaviour of an observed variable over a series of nested scales [61,62]. The brain's functional and corresponding dynamical heterogeneity produces a spatial distribution of time scales. While only some of these are usually considered of interest, global temporal scales may not coincide with any of the local dynamics, and may emerge from transient connectivity patterns created and destroyed by rewiring processes [63]. Because spatial scales also have non- trivial topological properties [64], global dynamics is a field endowed with arbitrarily real and phase-space complex topologies [65], where scales replace metric distances, fractal geometry the Euclidean one, and scale invariance Galilean invariance. By resorting to nonlinear analysis, algebraic and differential topology, renormalization group methods etc. [66,67], brain activity can then be described in terms of universal properties, i.e. regimes with robust macroscopic behaviour with respect to the nature of microscopic interactions and sharing the same symmetries. These descriptions partition the phase space, identify dynamical pathways leading to specific regions of this space, and allow relating descriptions of the same brain at different scales, and grouping descriptions of different brains exhibiting the same large-scale behaviour [66]. Chunking Correlated noise and cross-scale relationships produce temporally ordered structures which can help segmenting reasoning episodes into chunks. This can be done by defining quasi-stationary segments boundaries [68]. The scaling properties of quasi-stationary segments' durations may help clarifying whether reasoning in its general form is merely a repetition of simple cycles seen in more controlled forms of reasoning, or is of a qualitatively different nature, and in any case, determining the time scales at which simpler cycles are reedited. The waiting-time distribution between steps defines an internal operational time, which may grow sub- or super-linearly with physical time [69]. Multiplicative cross-scale interactions bias the waiting-time distribution so that operational and physical times no longer coincide, and local probability densities become time-dependent and intermittent [70]. FROM SPONTANEOUS ACTIVITY TO REASONING Cognitive processes can be thought of as selections and orchestrations of cortical states already present in spontaneous activity [71,72]. Each process corresponds to a specific phase space cut, with its own topological properties and symmetries, and characteristic kinematics, memory, ageing properties, degree of ergodicity, and internal clock [33]. Reasoning may modulate not brain activity's frequency or amplitude but its functional form [33], e.g. by pushing it towards the basin of attraction of advantageous probability distributions: good reasoning could be tantamount to designing a driving noise function forcing the system’s stationary distribution to equal a target one. Cognitive demands may generally change the symmetries of brain activity. For instance, scaling regime modulations observed in associations with a reasoning task [73] may correspond to cross-overs between universality classes, reflecting dynamical transitions in the system’s behaviour [74]. APPRAISING REASONING interpreted fluctuations can be Brain in various interrelated ways that help evaluating the quality of reasoning, by using models of the function it fulfils and quantifying faces while performing it. the constraints the brain Metaphors for reasoning search process [77]. This in Reasoning, as other cognitive processes, e.g. memory recall [75,76], can be represented as a search process similar to that of animals foraging in an unknown can be environment characterized terms of random walks [78-81]. Importantly, random walk types can quantify the extent to which a given trajectory optimizes search, given the characteristics of the explored space and the resources available to the individual [81]. Such a characterisation would allow assessing in a context-specific way the quality of both the reasoning and the 'reasoner'. That behavioural aspects of human cognition [75,76] and brain activity heavy-tailed distributions might indicate search optimality [80,82]. However, because these properties are generic in spontaneous activity, reasoning's quality can only be described in terms of its modulations. Furthermore, since local dynamics may be nowhere Lévy-like, finding the neural property and spatial scale showing scaling are the crucial steps. non-Gaussian, show both The reasoning regime could also be represented as a network traffic regulation problem, where phenomena such as overload or jamming may be quantified in terms of information creation, erasure and transmission rates, by regarding simple fluctuations as letters of an alphabet and fluctuation complexes as words, and quantifying the amount of information in the system. Characterizing traffic regulation may involve understanding the interplay between the underlying network's topology, burstiness of information packets and fluctuation distributions [83-85]. Although only causal information [86] may directly serve reasoning purposes, the total the shape of l correlates of reasoning Papo Characterizing the neura Papo Characterizing the neural correlates of reasoning l correlates of reasoning Papo Characterizing the neura Papo Characterizing the neura l correlates of reasoning information encoded in the network may describe the noise-control mechanisms indirectly optimizing it. The sudden onset of insight may be thought of as an extreme event comparable to earthquakes, financial crashes or epileptic seizures [87,88], e.g. as a rupture phenomenon, and the route to it as a long charging process, with nested hierarchical "earthquakes", and try predicting its occurrence. For such phenomena, the coupling strengths distribution and topology constitute the relevant field [89], which may be described using complex network theory [64]. It is tempting to conjecture that insight onset may be predicted by monitoring e.g. anomalous diffusion parameters [88], variations in Gaussianity fractal spectrum complexity [91,92]. [90], or changes in From dynamics to thermodynamics statistical aspects of brain Differences in reasoning abilities can produce not only differences in processing times of some orders of magnitude, but also qualitative ones in dynamical fluctuations, and and topology. in A is "How efficiently does a given brain carry then: out reasoning?". corresponding brain natural question when studying topography and reasoning There are various ways to assess efficiency of a given device. If reasoning was the output of an engine, e.g. a Carnot's engine, it would be natural to understand not only how the engine produces it, but also how efficiently it performs it. While functional network reconstruction can provide an indirect characterization of brain activity's energetic aspects [93,94], the brain's thermodynamics can directly be deduced from its dynamics [95], which can be interpreted as the walker's diffusion on the surface of the entropy production rate in the space of kinetic variables [96] . autocorrelation function in Temperature represents a good example of how thermodynamical variables can be used to describe brain activity [97]. For a system at equilibrium, temperature is proportional to the ratio between the response to an external field conjugate to some observable and the corresponding the unperturbed system [98]. In the brain [99], equilibrium temperature must be substituted by an effective temperature thermometer responding on the time scale at which the system reverts to equilibrium would measure [101]. Each scales can have its own effective temperature even within the same spatial region, corresponding to different distances from equilibrium and reflecting qualitatively different diffusion processes [102]. Thus, measuring effective temperature at various scales allows understanding the extent to which each from equilibrium, produces entropy, etc.. [100] reflecting what a spatio-temporal scale deviates Thermodynamic functions can be used to directly describe brain activity, but also as control parameters, i.e. one can monitor different aspects of brain activity as one measured thermodynamical variable varies in time. For instance, one may observe temperature variations during a reasoning task, but also possible phase transitions in some other property of neural activity, as temperature is varied [97]. FROM THEORY TO EXPERIMENT OBSERVING REASONING Reasoning is a difficult phenomenon to observe: tasks can be executed in more than one way, each possibly corresponding to a neural phase space with convoluted geometry and the processes involved in reasoning may evolve over time-scales exceeding those typical of laboratory testing. Proper observation of a given process requires that the observation time be much larger than any scale in the system. A process is observable if it has a finite ratio between the characteristic time of the independent variable and the length of the available time series [103]. Factors including long-term memory, aging and weak ergodicity breaking may result in a diverging ratio [104]. The observation time should also be much larger than the time needed to visit the neural phase space. However, deciding whether a reasoning episode, or even an ensemble of episodes, sufficiently sample a subject's repertoire is non-trivial. Cognitive neuroscientists observe phenomena through experiments where subjects typically carry out given tasks a large number of times, assumed to be independent realizations of the same observable, and to adequately sample the phase space of task-related brain activity. However, in the presence of complex fluctuations, trials may not self-average, i.e. dispersion would not vanish even for an infinite number of trials [105]. Furthermore, the time needed to explore this space may far exceed the typical reasoning episode duration, and reasoning episodes may explore different aspects of the space of available strategies. Thus, trials may improve phase space exploration rather than the signal-to-noise ratio [106]. EXPERIMENTAL IMPLICATIONS its Reasoning's characteristics, particularly lack of characteristic temporal duration, have implications at various levels. First, episodes cannot be compared in an event-related fashion. Second, defining reliable neural correlates of reasoning requires defining its characteristic temporal scales. Third, measures of brain activity should be invariant with respect to overall duration. Scaling exponents, data collapse and universality of fluctuations statistics [107-109], or explicit evolution equations for the particle's momenta and for the cross-scale fluctuation probabilities [62] can be retrieved from data and applied to unevenly lengthen trials. Thermodynamic quantities such as free energy or temperature can also be estimated for stochastic trajectories over finite time durations [97,110-113]. In all cases, the reconstruction of the Reasoning presents a dilemma between ensuring complete phase space exploration, which may require extremely long trials, and signal stationarity, which is guaranteed only for time scales much shorter than the reasoning episodes' duration. At fast time scales, the window in which relevant quantities are calculated should not introduce spurious time scales, filtering out genuine ones. Altogether, reasoning's inherently unstable nature suggests that describing it may boil down to characterizing non-stationarities and their aetiologies. Reasoning tasks may be so difficult that only few participants manage to produce solutions within a reasonable time. This represents a shortcoming when trials are considered as independent and identically distributed, as the signal-to-noise ratio improves with the square root of the number of trials. Smoothing response times is a frequent strategy to obviate this problem, but limits or distorts the reasoning process. Furthermore, however many, short trials may insufficiently explore the phase space. Designs with few long trials may express richer spatiotemporal brain dynamics than many short ones of equivalent overall length. Finally, while observed scaling properties may help understanding whether insight is predictable, i.e. whether it is an outlier or it is generated by the same distribution producing anonymous events, predicting insight onset in real data appears a challenging task, as reasoning episodes are various orders of magnitude shorter than earthquake, financial or epilepsy time series [114]. CONCLUSIONS Reasoning elicits an exceptionally rich repertoire of otherwise unexpressed neural properties. Its neural correlates to neuroscientists, whom it compels to consider hitherto neglected brain properties, as they are to psychologists striving to understand its underlying processes. as much helpful therefore are tools (borrowed Defining general and robust mechanistic properties of healthy and dysfunctional reasoning will require as yet non-standard brain metrics, experimental designs, and analytical rarely made available to psychologists). This may shed light fundamental mechanistic properties of both on and dysfunctional reasoning, and ultimately healthy help cognitive used and as brain to achieve desired states [115]. understanding pharmacological enhancers and interventions them targeting actions fields from the of ACKNOWLEDGEMENT author The (FIS201238949-C03-01). acknowledges 1. 2. 3. 4. 5. 6. 7. 8. 9. Moshman, D. (1995). Reasoning as self-constrained thinking. Hum. Devel. 38, 53–64. Gloning, K., and Hoff, H. (1969). Cerebral localization of disorders of higher nervous activity. In Vincken and Bruyn (Eds.), Handbook of clinical neurology (Vol. 3, Disorders of higher nervous activity). New York: Wiley. Lezak, M.D. (1995). Neuropsychological assessment. 3rd Edition. Oxford: Oxford University Press. Goel, V., Gold, B., Kapur, S., and Houle S. (1997). The seats of reason? An imaging study of deductive and inductive reasoning. Neuroreport 8, 1305–1310. Goel, V., Gold, B., Kapur, S., and Houle, S. (1998). Neuroanatomical correlates of human reasoning. J. Cogn. Neurosci. 10, 293–302. Noveck, I.A., Goel, V., and Smith, K.W. (2004). The neural basis of conditional reasoning with arbitrary content. Cortex 40, 613–622. Osherson, D., Perani, D., Cappa, S., Schnur, T., Grassi, F., and Fazio, F. (1998). Distinct brain foci in deductive versus probabilistic reasoning. Neuropsychologia 36, 369–376. Parsons, L.M., and Osherson, D. (2001). New evidence for distinct right and left brain systems for deductive versus probabilistic reasoning. Cereb. Cortex 11, 954–965. Prado, J., Spotorno, N., Koun, E., Hewitt, E., Van der Henst, J.B., Sperber, D., and Noveck, I.A. (2014). Neural interaction between logical reasoning and pragmatic processing in narrative discourse. J. Cogn. Neurosci. 16, 1–13. 10. Prado, J., Chadha, A., and Booth, J.R. (2011). The brain network for deductive reasoning: a quantitative meta- analysis of 28 neuroimaging studies. J. Cogn. Neurosci. 23, 3483–3497. 12. 11. Acuna, B.D., Eliassen, J.C., Donoghue, J.P., and Sanes, J.N. (2002). Frontal and parietal lobe activation during transitive inference in humans. Cereb. Cortex 12, 1312– 1321. Houdé, O., Zago, L., Crivello, F., Moutier, S., Pineau, A., Mazoyer, B., and Tzourio- Mazoyer, N. (2001). Access to deductive logic depends on a right ventromedial prefrontal area devoted to emotion and feeling: evidence from a training paradigm. Neuroimage 14, 1486– 1492. Kroger, J.K., Sabb, F.W., Fales, C.L., Bookheimer, S.Y., Cohen, M.S., and Holyoak, K.J. (2002). Recruitment of anterior dorsolateral prefrontal cortex in human reasoning: A parametric study of relational complexity. Cereb. Cortex 12, 477–485. 13. 14. Reverberi, C., Bonatti, L.L., Frackowiak, R.S., Paulesu, E., Cherubini, P., and Macaluso, E. (2012). Large scale brain activations predict reasoning profiles. Neuroimage 59, 1752–1764. P.-M. (2007). Baudonnière, 15. Papo, D., Douiri, A., Bouchet, F., Bourzeix, J.-C., Caverni, J.-P., and Time-frequency intracranial source localization of feedback-related EEG activity in hypothesis testing. Cereb. Cortex 17, 1314–1322. Bonnefond, M., Noveck I., Baillet S., Cheylus, A., Delpuech, C., Bertrand, O., Fourneret, P. and Van Der Henst, J.B. (2013). What MEG can reveal about reasoning: the case of if…then sentences. Hum. Brain Mapp. 34, 684–697. 16. l correlates of reasoning Papo Characterizing the neura Papo Characterizing the neural correlates of reasoning l correlates of reasoning Papo Characterizing the neura Papo Characterizing the neura l correlates of reasoning underlying dynamics device's resolution. improves with the recording REFERENCES the support of MINECO 17. Bonnefond, M., Castelain, T., Cheylus, A., and Van der Henst, J.B. (2014). Reasoning from transitive premises: an EEG study. Brain Cogn. 90, 100– 108. 18. Bonnefond, M., Kaliuzhna, M., Van der Henst, J.B., and De Neys, W. (2014). Disabling conditional inferences: an EEG study. Neuropsychologia 56, 255–262. l correlates of reasoning Papo Characterizing the neura Papo Characterizing the neural correlates of reasoning l correlates of reasoning Papo Characterizing the neura Papo Characterizing the neura l correlates of reasoning 19. Luo, J., Niki, K., and Phillips, S. (2004). Neural correlates of 20. 21. 22. 23. 24. the ‘Aha! reaction’. Neuroreport 15, 2013–2017. Subramaniam, K., Kounios, J., Parrish, T.B., and Jung- Beeman, M. (2008). A brain mechanism for facilitation of insight by positive affect. J. Cogn. Neurosci. 21, 415–432. Bowden, E.M., and Jung-Beeman, M. (2007). Methods for investigating the neural components of insight. Methods 42, 87–99. Jung-Beeman, M., Bowden, E.M., Haberman., J., Frymiare, J.L., Arambel-Liu, S., Greenblatt, R., Reber, P.J., and Kounios, J. (2004). Neural activity when people solve verbal problems with insight. PLoS Biol. 2, E97. Mai, X-Q., Luo, J., Wu, J-H., and Luo, Y-J. (2004). “Aha!” Effects in a guessing riddle task: An event-related potential study. Hum. Brain Mapp. 22, 261–270. Lang, S., Kanngieser, N., Jaśkowski, P., Haider, H., Rose, M., and Verleger, R. (2006). Precursors of insight in event- related brain potentials. J. Cogn. Neurosci. 18, 2052–2066. 39. Dragoi, G., and Tonegawa, S. (2011 ). Preplay of future place cell sequences by hippocampal cellular assemblies. Nature 469, 397–401. 40. Betzel, R.F., Erickson, M.A., Abell, M., O'Donnell, B.F., Hetrick, W.P., and Sporns, O. (2012). Synchronization dynamics and evidence for a repertoire of network states in resting EEG. Front. Comput. Neurosci. 6, 74. 41. Gong, P., Nikolaev, A.R., and van Leeuwen, C. (2007). Intermittent dynamics underlying the intrinsic fluctuations of the collective synchronization patterns in electrocortical activity. Phys. Rev. E 76, 011904. 42. Allegrini, P., Menicucci, D., Paradisi, P., and Gemignani, A. (2010). Fractal spontaneous EEG metastable-state transitions: new vistas on integrated neural dynamics. Front. Physio. 1, 128. complexity in 43. Bianco, S., Ignaccolo, M., Rider, M.S., Ross, M.J., Winsor, P., and Grigolini, P. (2007). Brain, music, and non-Poisson renewal processes. Phys. Rev. E 75, 061911. 25. Qiu, J., Li, H., Yang, D., Luo, Y., Li, Y., Wu, Z., and Zhang, Q. (2008). The neural basis of insight problem solving: an event-related potential study. Brain Cogn. 68, 100–106. 44. Novikov, E., Novikov, A., Shannahoff-Khalsa, D., Schwartz, B., and Wright, J. (1997). Scale-similar activity in the brain. Phys. Rev. E 56, R2387–R2389. 27. 26. Kounios, J., Fleck, J.I., Green, D.L., Payne, L., Stevenson, J.L., Bowden, E.M., and Jung- Beeman, M. (2008). The origins of insight in resting-state brain activity. Neuropsychologia 46, 281–291. Kounios, J., Frymiare, J.L., Bowden, E.M., Fleck, J.I., Subramaniam, K., Parrish, T.B., and Jung-Beeman, M. (2006). The prepared mind: neural activity prior to problem presentation predicts subsequent solution by sudden insight. Psychol. Sci. 17, 882–890. 28. Sheth, B.R., Sandkühler, S., and Bhattacharya, J. (2008). Posterior beta and anterior gamma oscillations predict cognitive insight. J. Cogn. Neurosci. 21, 1269–1279. 29. Sandkühler, S., and Bhattacharya, J. (2008). Deconstructing insight: EEG correlates of insightful problem solving. PLoS ONE 3, e1459. 30. Knoblich, G., Ohlsson, S., Haider, H., and Rhenius, D. (1999). Constraint relaxation and chunk decomposition in insight problem solving. J. Exp. Psychol. Learn. 25, 1534–1556. 31. Stephen, D.G., Boncoddo, R. A., Magnuson, J. S., & Dixon, J. A. (2009). The dynamics of insight: mathematical discovery as a phase transition. Mem. Cognition 37, 1132–1149. 32. Rabinovich, M., Huerta, R., and Laurent, G. (2008). Transient dynamics for neural processing. Science 321, 48–50. 33. Papo (2014). Functional complex in brain activity: from resting state to significance of fluctuations cognitive neuroscience. Front. Syst. Neurosci. 8, 112. 34. Kenet, T., Bibitchkov, D., Tsodyks, M., Grinvald, A., and Arieli, A. (2003). Spontaneously emerging cortical representations of visual attributes. Nature 425, 954–956. 35. Cossart, R., Aronov, D., and Yuste, R. (2003). Attractor dynamics of network UP states in the neocortex. Nature 423, 283–288. Ikegaya, Y., Aaron, G., Cossart, R., Aronov, D., Lampl, I., Ferster, D. and Yuste, R. (2004). Synfire chains and cortical songs: temporal modules of cortical activity. Science 304, 559–564. 36. 37. Beggs, J.M., and Plenz, D. (2003). Neuronal avalanches in neocortical circuits. J. Neurosci. 23, 11167–11177. 38. Beggs, J.M., and Plenz, D. (2004). Neuronal avalanches are diverse and precise activity patterns that are stable for many hours in cortical slice cultures. J. Neurosci. 24, 5216– 5229. 45. Linkenkaer-Hansen, K., Nikouline, V.V., Palva, J.M., and Ilmoniemi, R. (2001). Long-range temporal correlations and scaling behavior in human oscillations. J. Neurosci. 15, 1370–1377. 46. Gong, P., Nikolaev, A.R., and van Leeuwen, C. (2002). Scale- invariant fluctuations of the dynamical synchronization in human brain electrical activity. Neurosci. Lett. 336, 33–36. 47. Stam, C.J., and de Bruin, E.A. (2004). Scale-free dynamics of global functional connectivity in the human brain. Hum. Brain Mapp. 22, 97–109. 48. Freeman, W.J., Holmes, M.D., Burke, B.C., and Vanhatalo, S. (2003). Spatial spectra of scalp EEG and EMG from awake humans. Clin. Neurophysiol. 114, 1053–1068. 49. van de Ville, D., Britz, J., and Michel, C.M. (2010). EEG microstate sequences in healthy humans at rest reveal scale-free dynamics. Proc. Natl. Acad. Sci. USA 107, 18179– 18184. 50. Expert, P., Lambiotte, R., Chialvo, D.R., Christensen, K., Jensen, H.J., Sharp, D.J., and Turkheimer, F. (2010). Self- similar correlation resting-state functional magnetic resonance imaging. J. R. Soc. Interface 8, 472–479. in brain function 52. 51. Fraiman, D., and Chialvo D.R. (2012). What kind of noise is brain noise: anomalous scaling behavior of the resting brain activity fluctuations. Front. Physiol. 3, 307. Suckling, J., Wink, A.M., Bernard, F.A., Barnes, A., and Bullmore, E. (2009). Endogenous multifractal brain dynamics are modulated by age, cholinergic blockade and cognitive performance. J. Neurosci. Methods 174, 292– 300. 53. Zilber, N., Ciuciu, P., Abry, P., and van Wassenhove, V. (2012). Modulation of scale-free properties of brain activity in MEG. IEEE I. S. Biomed. Imaging (Barcelona), 1531–1534. 54. Freyer, F., Aquino, K., Robinson, P.A., Ritter, P., and Breakspear, M. (2009). Bistability and Non-Gaussian fluctuations in spontaneous cortical activity. J. Neurosci. 29, 8512–8524. 55. Hsu, D., and Hsu, M. (2009). Zwanzig-Mori projection operators and EEG dynamics : Deriving a simple equation of motion. PMC Biophysics 2, 6. l correlates of reasoning Papo Characterizing the neura Papo Characterizing the neural correlates of reasoning l correlates of reasoning Papo Characterizing the neura Papo Characterizing the neura l correlates of reasoning 56. Grigolini, P., Rocco, A., and West, B.J. (1999). Fractional calculus as a macroscopic manifestation of randomness. Phys. Rev. E 59, 2603–2613. 76. Baronchelli, A., and Radicchi, F. (2013). Lévy flights in human behaviour and cognition. Chaos Soliton. Fract. 56, 101–105. 57. Papo, D. (2013). Time scales in cognitive neuroscience. Front. Physio. 4, 86. 58. Fairhall, A.L., Lewen, G.D., Bialek, W., and de Ruyter van Steveninck, R. (2001). Multiple timescales of adaptation in a neural in Neural Information Processing Systems 13. ( Leen TK, Dietterich TG, Tresp V. Eds.), MIT Press, pp. 124–130. In Advances code. 59. Gilboa, G., Chen, R., and Brenner N. (2005). History- dependent multiple-timescale dynamics in a single-neuron model. J. Neurosci. 25, 6479–6489. 60. Lundstrom, B.N., Higgs, M.H., Spain, W.J., and Fairhall, A.L. neocortical (2008). pyramidal neurons. Nat. Neurosci. 11, 1335–13342. differentiation Fractional by 61. Bacry, E., Delour, J., and Muzy, J.F. (2001). Multifractal random walk. Phys. Rev. E 64, 026103. 62. Friedrich, R., Peinke, J., Sahimi, M., and Reza Rahimi Tabar, M. (2011). Approaching complexity by stochastic methods: from biological systems to turbulence. Phys. Rep. 506, 87– 162. 63. Bianco, S., Geneston, E., Grigolini, P., and Ignaccolo, M. (2007). Renewal aging as emerging property of phase synchronization. Physica A 387, 1387–1392. 77. Viswanathan, G.M., da Luz, M.G.E., Raposo, E.P., and Stanley, H.E. (2011). The physics of foraging: an introduction to random searches and biological encounters. Cambridge University Press, Cambridge. 78. Shlesinger, M., Zaslavsky, G., and Klafter, J. (1993). Strange kinetics. Nature 363, 31–37. 79. Codling, E.A., Plank, M.J., and Benhamou, S. (2008). Random walk models in biology. J. R. Soc. Interface 5, 813- 834. in 82. intermittent search processes 80. Lomholt, M., Tal, K., Metzler, R., and Joseph, K. (2008). Lévy strategies are advantageous. Proc. Natl. Acad. Sci. USA 105, 11055–11059. 81. Bénichou, O., Loverdo, C., Moreau, M., and Voituriez, R. (2011). Intermittent search strategies. Rev. Mod. Phys. 83, 81. Humphries, N.E., Weimerskirch, H., Queiroz, N., Southall, E.J., and Sims, D.W. (2012). Foraging success of biological Lévy flights recorded in situ. Proc. Natl. Acad. Sci. USA 109, 7169–7174. DeDeo, S., and Krakauer, D.C. and Soc. Interface 9, 2131–2144. (2012). Dynamics in finite self-similar networks. J. R. processing 83. 64. Bullmore, E.T., and Sporns, O. (2009). Complex brain networks: graph theoretical analysis of structural and functional systems. Nat. Rev. Neurosci. 10, 186–198. 84. Delvenne, J-C., Lambiotte, R., and Rocha, L.E.C. (2013). Bottlenecks, burstiness, and fat tails regulate mixing times of non-Poissonian random walks. arXiv:1309.4155. 65. Zaslavsky, GM. (2002). Chaos, fractional kinetics, and anomalous transport. Phys. Rep. 371, 461–580. 66. Lesne, A. (2008). Regularization, renormalization, and renormalization groups: relationships and epistemological aspects. In. Vision of Oneness. (I. Licata and A. Sakaji Eds), Aracne, Roma, pp. 121–154. 67. Petri, G., Expert, P., Turkheimer, F., Carhart-Harris, R., Nutt, D., Hellyer, J., and Vaccarino, F. (2014). Homological scaffolds of brain functional networks. J. R. Soc. Interface 11, 20140873. 68. Kaplan, A.Y., Fingelkurts, A.A., Fingelkurts, A.A., Borisov, B.S., and Darkhovsky, B.S. (2005). Nonstationary nature of the brain activity as revealed by EEG/MEG: methodological, practical and conceptual challenges. Signal Process. 85, 2190– 2212. Sokolov, I.M., and Klafter, J. (2005). From diffusion to anomalous diffusion: a century after Einstein’s Brownian motion. Chaos 15, 026103. 69. 70. Crisanti, A., and Ritort, F. (2003). Violation of the fluctuation–dissipation theorem in glassy systems: basic notions and the numerical evidence. J. Phys. A-Math. Gen. 36, R181–R290. 71. Fiser, J., Chiu, C., and Weliky, M. (2004). Small modulation of ongoing cortical dynamics by sensory input during natural vision. Nature 431, 573–578. 72. Luczak, A., Barthó, P., and Harris, K.D. (2009). Spontaneous events outline the realm of possible sensory responses in neocortical populations. Neuron 62, 413–425. 73. Buiatti, M., Papo, D., Baudonnière, P.M., and van Vreeswijk, C. (2007). Feedback modulates the temporal scale-free dynamics of brain electrical activity in a hypothesis testing task. Neuroscience 146, 1400–1412. 74. Burov, S., and Barkai, E. (2008). Critical exponent of the fractional Langevin equation. Phys. Rev. Lett. 100, 070601. 75. Rhodes, T., and Turvey, M.T. (2007). Human memory retrieval as Lévy foraging. Physica A 385, 255–260. 85. Lambiotte, R., Tabourier, L., and Delvenne, J-C. (2013). Burstiness and spreading on temporal networks. Eur. Phys. J. B 86, 320. 86. Shalizi, C.R., and Moore, C. (2003). What is a macrostate? objective dynamics. and Subjective arXiv:cond-mat/0303625v1. observations 87. Osorio, I., Frei, MG, Sornette, D, Milton, J, and Lai YC. (2010 ). Epileptic seizures: quakes of the brain? Phys. Rev. E 82, 021919. 88. Contoyiannis Y.F., and Eftaxias, K.A. (2008 ). Tsallis and Levy statistics in the preparation of an earthquake. Nonlin. Processes Geophys. 15, 379–388. 89. Sornette, D., and Osorio, I. (2010 ). Prediction. arXiv:1007.2420v1. 90. Manshour, P., Saberi, S., Sahimi, M. Peinke, J., Pacheco, A.F., and Rahimi Tabar, M.R. (2009). Turbulencelike behavior of seismic time series. Phys. Rev. Lett. 102, 014101. 91. de Arcangelis, L., and Herrmann, H.J. (1989). Scaling and multiscaling laws in random fuse networks. Phys. Rev. B 39, 2678. 92. Kapiris, P.G., Eftaxias, K.A., and Chelidze, T.L. (2004). Electromagnetic signature of prefracture criticality in heterogeneous media. Phys. Rev. Lett. 92, 065702. 93. Bullmore, E.T., and Sporns, O. (2012). The economy of brain network organization. Nat. Rev. Neurosci. 13, 336- 349. 94. Papo, D., Zanin, M., Pineda, J.A., Boccaletti, S., & Buldú, J.M. (2014). Brain networks: great expectations, hard times, and the big leap forward. Phil. Trans. R. Soc. B 369, 20130525. 95. Sekimoto, and thermodynamics. Prog. Theor. Phys. Supplement 130, 17– 27. Langevin equation (1998). K. 96. Kirkaldy, J.S. (1964 ). Thermodynamics of the human brain. Biophys. J. 5, 981–986. l correlates of reasoning Papo Characterizing the neura Papo Characterizing the neural correlates of reasoning l correlates of reasoning Papo Characterizing the neura Papo Characterizing the neura l correlates of reasoning 97. Papo, D. (2013). Brain temperature: what it means and what neuroscientists. arXiv:1310.2906v1. (cognitive) can do for it 98. Kubo, R. (1966). The fluctuation-dissipation theorem. Rep. 99. Progr. Phys. 29, 255–284. Papo, D. (2013). Why should cognitive neuroscientists study the brain's resting state? Front. Hum. Neurosci. 7, 45. doi: 10.3389/fnhum.2013.00045 100. Cugliandolo, L. F. (2011). The effective temperature. J. Phys. A Math. Theor. 44, 483001. 101. Kurchan, J. (2005). In and out of equilibrium. Nature 433, 222–225. 102. Papo, D. (2014). Measuring brain temperature without a thermometer. Front. Physio. 5, 124. 103. Reiner, M. (1964). The Deborah number. Phys. Today 17, 62. 104. Rebenshtok, A., and Barkai, E. (2007). Distribution of time- averaged observables for weak ergodicity breaking. Phys. Rev. Lett. 99, 210601. 105. Aharony, A., and Harris, A.B. (1996). Absence of self- averaging and universal fluctuations in random systems near critical points. Phys. Rev. Lett. 77, 3700–3703. 106. Ghosh, A., Rho, Y., McIntosh, A.R., Kötter, R., and Jirsa, V.K. (2007). Noise during rest enables the exploration of the brain's dynamic repertoire. PLoS Comput. Biol. 4, e1000196. 107. Bramwell, S.T., Holdsworth, P.C.W., and Pinton, J.-F. (1998). in turbulence and Universality of rare critical phenomena. Nature 396, 552–554. fluctuations 108. Bhattacharya, J. (2009). Increase of universality in human brain during mental imagery from visual perception. PLoS One 4, e4121. 109. Friedman, N., Ito, S., Brinkman, B.A., Shimono, M., Deville, R.E., Dahmen, K.A., Beggs, J.M., and Butler, T.C. (2012). Universal critical dynamics in high resolution neuronal avalanche data. Phys. Rev. Lett. 108, 208102. Thermodynamic 110. Ruelle, formalism. (1978). D. Addison Wesley Publ. Co., Reading, MA. 111. Beck, C., and Schlögl, F. (1997). Thermodynamics of chaotic systems: an introduction. Cambridge University press, Cambridge. 112. Canessa, E. (2000). Multifractality in time series. J. Phys. A: Math. Gen. 33, 3637–3651. 113. Olemskoi, A., and Kokhan, S. (2006). Effective temperature of self-similar time series: analytical and numerical developments. Physica A 360, 37–58. 114. Sornette, D. (2002). Predictability of catastrophic events: Material rupture, earthquakes, financial crashes, and human birth. Proc. Natl. Acad. Sci. USA 99, 2522–2529. turbulence, 115. Gutiérrez R, Sendiña-Nadal I, Zanin M, Papo D, Boccaletti S. 2012 Targeting the dynamics of complex networks. Sci. Rep. 2, 396.
1705.04739
1
1705
2017-05-12T20:02:11
Face recognition assessments used in the study of super-recognisers
[ "q-bio.NC" ]
The purpose of this paper is to provide a brief overview of nine assessments of face processing skills. These tests have been used commonly in recent years to gauge the skills of perspective 'super-recognisers' with respect to the general population. In the literature, a person has been considered to be a 'super-recogniser' based on superior scores on one or more of these tests (cf., Noyes, Phillips & O'Toole, in press). The paper provides a supplement to a recent review of super-recognisers aimed at readers who are unfamiliar with these tests. That review provides a complete summary of the super-recongiser literature to date (2017). It also provides a theory and a set of action points directed at answering the question "What is a super-recogniser?"
q-bio.NC
q-bio
FACE RECOGNITION ASSESSMENTS USED IN THE STUDY OF SUPER-RECOGNISERS Eilidh Noyes*,1 and Alice J. O'Toole1 1 School of Behavioural and Brain Sciences, The University of Texas at Dallas, Richardson, USA The purpose of this paper is to provide a brief overview of nine assessments of face processing skills. These tests have been used commonly in recent years to gauge the skills of perspective 'super-recognisers' with respect to the general population. In the literature, a person has been considered to be a 'super- recogniser' based on superior scores on one or more of these tests (cf., Noyes, Phillips, & O'Toole, in press). The paper provides a supplement to a recent review of super-recognisers aimed at readers who are unfamiliar with these tests. That review provides a complete summary of the super-recoginser literature to date (2017). It also provides a theory and a set of action points directed at answering the question "What is a super-recogniser?" Before They Were Famous Test (BTWFT) The Before They Were Famous Task (BTWFT) was designed to test accuracy in recognising famous faces across age-related changes. The task uses 56 images of famous identities, taken at a time before each celebrity was famous (usually a photograph of the celebrity that had been taken during childhood). Participants are asked to identify each of the faces. A correct identification is accepted if the participant correctly names the face, or if they provide a correct uniquely identifying fact for the celebrity. Results for this task are calculated in terms of the percentage of correct responses. The Cambridge Face Memory Task Long Form (CFMT Long-Form) (Russell, Duchaine, & Nakayama, 2009) The CFMT long-form is an extension of the CFMT (Duchaine & Nakayama, 2006), a task originally used to distinguish prosopagnosics from controls. The original CFMT task requires participants to learn six unfamiliar male faces by viewing them from different viewpoints during a learning phase. Participants are tested subsequently in three different test conditions, each * Corresponding Author address Email: [email protected] 2 Eilidh Noyes and Alice J. O'Toole progressively more difficult than the previous one. First, participants are tested on their ability to recognise the exact image that they learned of each identity. Next, they are tested on whether they can recognise the faces in new images. In the third test, they must recognise the identities in novel images to which noise has been added. Across the three tests, there are a total of 72 recognition trials. After testing just one super-recogniser on the task, it was clear to Russell et al. (2009) that the CFMT was not suitable to test super-recognisers, due to ceiling levels of performance. Therefore, a fourth section was added to create the CFMT long-form. This section consisted of very challenging trials-the test images of previously learned faces varied drastically from the images that were learned. The images used at training were all of good quality, and had been cropped to include only the face. Test images in the fourth section were not cropped and had added extra noise (more than in section three), or were profile images with added extra noise. To make the task even more challenging, distractor test images (image of faces that had not been learned) were repeated more often than any of the previous conditions. This increased familiarity with the distractor items made them harder to distinguish from other known (i.e., previously learned) identities. Comparison of performance (percentage accuracy) across the sections of the test confirmed that section four of CFMT long form was the most difficult of the sections (Russell et al., 2009). Cambridge Face Perception Test (CFPT) (Duchaine, Germine, & Nakayama, 2007) The CFPT was designed to assess face perception ability – i.e., identification of faces without a memory component. Participants are presented with a target face image (in ¾ view) and frontal images of morphs between the target and six other faces. These images were created by morphing the target face with the faces of six other individuals. The morph proportion between the individual and target faces retains differing degrees of similarity (e.g., 88%, 76%, 64%, 52% 28%) to the target face. Participants rank the six faces from one to six, in order of similarity with the target face. The correct rank order would place the face with the greatest proportion of the target image as most similar to the target face, followed by the image with the next greatest percentage of the target face and so on. Participants are tested with both upright and inverted images (8 trial types of each of these conditions). This test design makes it possible to measure face matching through similarity ratings. The inclusion of upside-down faces also allows for testing of an inversion effect. What is a Super-Recogniser? 3 The Glasgow Face Matching Task (short version) (Burton, White, & McNeill 2010) The Glasgow Face Matching Task (GFMT) (short version) is a standardised test of unfamiliar face matching with well-established general population norms. The task consists of the most difficult 40 trials from the Glasgow Face Matching Task (Burton, White, & McNeill, 2010). When completing the GFMT short form, participants view face image pairs and are tasked with deciding whether the images in each pair are two different images of the same identity or are instead images of two different people (see Figure 1). Results for this test are calculated in terms of the percentage of correct responses made. Figure 1. Example of a same- (top) and different-identity (bottom) trial from the GFMT. Figure image from Robertson et al. (2016). Models Face Matching Test (MFMT) (Dowsett & Burton, 2014) The Models Face Matching Test (MFMT) is similar in design to the GFMT, but consists of more challenging face stimuli. The stimuli in the models face matching test are images of models, with each image captured from a different photoshoot. These images capture large differences in appearance across multiple images of the same identity. In the MFMT participants view face image pairs and make same or different identity decisions for each image pair. As in the GFMT, this test uses percentage accuracy measures of performance. 4 Eilidh Noyes and Alice J. O'Toole The Pixelated Lookalike Test (Noyes & Jenkins, 2017) The Pixelated Lookalike Test (PLT) is a challenging face matching task designed to assess the effect of familiarity on face matching performance for individuals of very similar appearance presented in poor quality images (Noyes & Jenkins, 2017). The task emulates that of a standard face matching test, with participants making same or different identity judgements to image pairs. Match trials in this task consist of two different images of a celebrity, whereas mismatch trials (different identity trials) consist of a celebrity image alongside an image of a professional celebrity 'lookalike' for that celebrity. This means that the people in the mismatch trials appear very similar, making this task challenging. The task is made even more difficult by presenting the images in pixelated form (see Figure 2). Figure 2. Example of the type image pairs used in the PLT. Image pair on the left shows different identities (with the imposter image on the right) and the pair on the left shows two images of the same identity. Figure from Robertson et al. (2016). Pixelation helps to mirror the image conditions that may be encountered in poor quality surveillance footage or digital imaging devices. Following completion of the face matching task, participants are measured on their familiarity with each of the celebrities in the task. This familiarity measure allows for performance accuracy to be broken down according to personal levels of familiarity with each face. Therefore, accuracy for each individual is known for faces that are completely unfamiliar, somewhat familiar, and extremely familiar. A percentage of correct responses is thus obtained for each participant at each familiarity band. What is a Super-Recogniser? 5 The Navon Test (Navon, 1977) The Navon test is used to provide a measure of processing bias for global compared to local image features (Navon, 1977). In the test, participants are presented with stimuli taking the appearance of an alphabetical letter, but the shape of this letter is formed by the configuration of many small letters. For example, participants may see a large 'H' on the screen but a close look at the 'H' may reveal that it is built up from many small 'T's. Participants are asked to report the large letter (in half of the trials) and the small letter (in the other half of the trials). Calculation of response error allows detection of bias to global or local features. Composite Face Test (Young, Hellawell & Hay, 1987) The composite face test is a test of holistic processing (Young, Hellawell, & Hay, 1987). Participants view composite and non-composite face images, in random order, and are asked to make identity judgments for either the top or bottom portion of the face (block dependant). A composite image consists of the top half of one face combined with the bottom half of another face, aligned into the configuration of a normal face. A non-composite image again consists of the top half of one face combined with the bottom half of another face, but misaligned horizontally. Naming accuracy is recorded and compared for the top section of the face presented in composite and non-composite form, and the bottom section of the face in composite and non-composite presentation format. Mean reaction time for accurate identification is generally lower for the non-composite image condition. This is explained by holistic processing-two identity halves presented together are 'combined' and perceived as a new identity. Famous Face Recognition Test (Lander et al. 2001) In the Famous Face Recognition Test participants are presented with 30 celebrity face images, as well as 10 unfamiliar faces (half of all images were male, the other female). Each of these images are presented one at a time and are viewed in either moving or static form (half of the images viewed in each of these conditions). The participants' task is to identify each face, by naming or providing a unique identifying fact about the face. If the face was an unknown (unfamiliar) face, the correct answer is to state that the face is unknown. After the task participant familiarity with the familiar faces is verified by providing them with a list of names of the faces that they had seen and asking whether they should have recognised the face of that celebrity. Percentage accuracy of response is calculated. 6 Eilidh Noyes and Alice J. O'Toole REFERENCES Burton, a M., White, D., & McNeill, A. (2010). The Glasgow Face Matching Test. Behavior Research Methods, 42, 286–91. Dowsett, A. J., & Burton, a. M. (2014). Unfamiliar face matching: Pairs out-perform individuals and provide a route to training. British Journal of Psychology, 106, 433–445. Duchaine, B., & Nakayama, K. (2006). The Cambridge Face Memory Test: Results for neurologically intact individuals and an investigation of its validity using inverted face stimuli and prosopagnosic participants. Neuropsychologia, 44, 576–585. Noyes, E., & Jenkins, R. (2017). Improving Performance on a Difficult Face Matching Task. Perception, 46, 226. Robertson, D. J., Noyes, E., Dowsett, A., Jenkins, R., & Burton, A. M. (2016). Face recognition by Metropolitan Police Super-recognisers. Plos One, 11, e0150036. Russell, R., Duchaine, B., & Nakayama, K. (2009). Super-recognizers: people with extraordinary face recognition ability. Psychonomic Bulletin & Review, 16, 252–257. Young, A., Hellawell, D., & Hay, D. (1987). Configural information in face perception. Perception, 16, 747–759.
1910.05271
1
1910
2019-10-08T19:33:49
A Test for Shared Patterns in Cross-modal Brain Activation Analysis
[ "q-bio.NC", "cs.LG", "stat.ML" ]
Determining the extent to which different cognitive modalities (understood here as the set of cognitive processes underlying the elaboration of a stimulus by the brain) rely on overlapping neural representations is a fundamental issue in cognitive neuroscience. In the last decade, the identification of shared activity patterns has been mostly framed as a supervised learning problem. For instance, a classifier is trained to discriminate categories (e.g. faces vs. houses) in modality I (e.g. perception) and tested on the same categories in modality II (e.g. imagery). This type of analysis is often referred to as cross-modal decoding. In this paper we take a different approach and instead formulate the problem of assessing shared patterns across modalities within the framework of statistical hypothesis testing. We propose both an appropriate test statistic and a scheme based on permutation testing to compute the significance of this test while making only minimal distributional assumption. We denote this test cross-modal permutation test (CMPT). We also provide empirical evidence on synthetic datasets that our approach has greater statistical power than the cross-modal decoding method while maintaining low Type I errors (rejecting a true null hypothesis). We compare both approaches on an fMRI dataset with three different cognitive modalities (perception, imagery, visual search). Finally, we show how CMPT can be combined with Searchlight analysis to explore spatial distribution of shared activity patterns.
q-bio.NC
q-bio
A Test for Shared Patterns in Cross-modal Brain Activation Analysis Elena Kalinina1,2,*, †, ‡, Fabian Pedregosa3, †, **, Vittorio Iacovella2, Emanuele Olivetti1,2, and Paolo Avesani1,2 1NeuroInformatics Laboratory (NILab), Bruno Kessler Foundation, Trento, Italy 2Centro Interdipartimentale Mente e Cervello (CIMeC), University of Trento, Italy 3Department of Electrical Engineering, UC Berkeley, California *e [email protected] †These authors contributed equally to the paper. ‡Work done while PhD student at NeuroInformatics Laboratory (NILab) and Centro Interdipartimentale Mente e Cervello (CIMeC). ** Work done while post-doctoral researcher at the Department of Electrical Engineering, UC Berkeley. ABSTRACT Determining the extent to which different cognitive modalities (understood here as the set of cognitive processes underlying the elaboration of a stimulus by the brain) rely on overlapping neural representations is a fundamental issue in cognitive neuroscience. In the last decade, the identification of shared activity patterns has been mostly framed as a supervised learning problem. For instance, a classifier is trained to discriminate categories (e.g. faces vs. houses) in modality I (e.g. perception) and tested on the same categories in modality II (e.g. imagery). This type of analysis is often referred to as cross-modal decoding. In this paper we take a different approach and instead formulate the problem of assessing shared patterns across modalities within the framework of statistical hypothesis testing. We propose both an appropriate test statistic and a scheme based on permutation testing to compute the significance of this test while making only minimal distributional assumption. We denote this test cross-modal permutation test (CMPT). We also provide empirical evidence on synthetic datasets that our approach has greater statistical power than the cross-modal decoding method while maintaining low Type I errors (rejecting a true null hypothesis). We compare both approaches on an fMRI dataset with three different cognitive modalities (perception, imagery, visual search). Finally, we show how CMPT can be combined with Searchlight analysis to explore spatial distribution of shared activity patterns. Introduction Functional MRI recordings enable the investigation of activation patterns that characterize the working brain. The main goal is to detect whether the neural pattern of a region of interest correlates with a cognitive task, like, for example, object category identification. Such investigations are usually focused on a specific cognitive modality, i.e.: visual perception, real auditory, visual imagery, auditory imagery, etc. A qualitative discrimination task can be designed to extract relevant information from patterns activated by two (or more) stimuli categories (like body and car) in one of these cognitive modalities. Identification of activation patterns that are shared across modalities has been the subject of numerous neurocognitive studies, with such modalities as mental calculations1, sensory/motor stimulation2 or words and picture-viewing3, to name a few. The most common approach here is to cast the problem of identifying common activation patterns as a supervised learning, or brain decoding, problem4, 5. Nastase and colleagues9 point out that successful classification in this setting allows to conclude that neural patterns elicited by relevant cognitive factors in one modality generalize accross to the patterns in the other modality. In other words, a low misclassification error on a modality which is different from the one used to train the classifier provides empirical evidence that a given region of interest is involved in the cognitive task encoded in both modalities. In the literature, such approach is referred to as cross-modal decoding analysis (CMDA). Statistical significance of its result can be assessed using a t-test on the accuracy obtained by the classifier1, 6, or by a permutation test based on computing the null distribution2, 7, 8 of such statistic. However, CMDA suffers from a number of practical issues. As we are going to discuss in section Methods, the accuracy of a decoding model is often low in a cross-modal setting, which most probably implies an exaggerated amount of Type II errors (failure to reject the null hypothesis). Neuroscientific investigations into the activity patterns in the fMRI data studies can be formulated as "confirmatory" or "exploratory" analysis. Confirmatory analysis is centered on a pre-established region of interest (ROI). The CMDA method presented above is often used for the confirmatory approach, when it is run on the data coming from a predefined ROI (or a set of predefined ROIs). Exploratory analysis aims at localization of areas containing information about the presented stimuli. Here, CMDA is employed in conjunction with Searchlight technique to explore the spatial structure of cross-modal activations9. The outcome of the Searchlight procedure are maps, where each voxel is assigned some quantitative measure of information that the voxel contains about the stimulus. The procedure of obtaining the maps consists in applying decoding sphere by sphere on time series extracted from the sphere voxels, and most commonly used information measure to produce Searchlight maps is classification accuracy. Maps are first calculated individually for each subject and then pooled together to run group analysis, where the significance of the obtained values is typically established running t-tests voxelwise with respect to chance level. Classification accuracy and t-tests have been subject to numerous criticisms with regard to their role in Searchlight analysis23. Besides, use of Searchlight for cross-modal analysis faces interpretation challenges introduced by the asymmetries both in accuracies and p-values when training and testing on data coming from two different cognitive modalities9. In this work we develop a permutation test for the investigation of shared patterns across modalities that we denote cross-modal permutation test (CMPT). This test builds on a long tradition of randomization inference in the statistics literature, which can be traced back to the first half of the 20th century10 -- 12. Permutation tests have recently seen renewed interest in neuroimaging13 -- 16 thanks to their minimal distributional assumptions and the availability of cheap computational resources. We provide empirical evidence on synthetic datasets that this method reduces Type II errors (failure to reject a false null hypothesis) while maintaining Type I errors comparable (incorrect rejection of the null hypothesis) with respect to CMDA. Our results highlight particular advantages of CMPT in the small sample/high dimensional regime, a setting of practical importance in neuroimaging studies. Next, we compare CMPT and CMDA on an fMRI study of three cognitive modalities: visual attention, imagery and perception. We conduct confirmatory analysis comparing the performance of CMPT and CMDA when identifying the presence of shared patterns within a functionally defined Region of Interest (ROI). Finally, we present the results of an exploratory analysis with Searchlight making use of the proposed CMPT test for the information based mapping to explore the presence of common patterns between different modalities at the whole brain level. The use of CMPT allows to overcome major methodological drawbacks that had been pointed out for Searchlight in the literature23. Methods Cross-modal permutation test (CMPT) In this section we describe a statistical test for cross-modal activation pattern analysis that we denote CMPT. We formulate the problem of assessing cross-modal activation as a hypothesis testing problem and propose an inference procedure for this test based on a permutation schema. Setting. We assume that the experimental task consists of two modalities (e.g., auditory and visual perception) and each image in the dataset containing an activation pattern has an associated condition A or B (e.g., two stimuli categories like human body and car). In total, we observe n activation patterns for each modality, corresponding to the number of conditions in the experiment, where each activation image is a mean image (averaged by the number of trials in the experiment) representative of one condition, and the goal is to decide whether there is a common condition effect across the different modalities. Let us formalize this in the language of statistical hypothesis testing. Consider the set of pairs Z(X,Y ) = {(X1,Y1), . . . , (Xn,Yn)} sampled iid from some unknown probability distribution P, where X = {X1, . . . ,Xn} (resp. Y = {Y1, . . . ,Yn}) are the activation patterns corresponding to the first (resp. second) modality, and where the experimental paradigm is designed such that the Xi and Yi are associated with the same condition (but different modality). Since the image pairs belong to the same condition, as long as there is a condition effect shared across modality, the sequences X and Y cannot be independent. We hence formulate the null hypothesis (which we want to reject) that both sequences are independent and so their joint probability distribution P factorizes over their marginal: H0 : P = PX × PY , with PX, PY the marginal distribution of P. (1) Test statistic. Given the set of image pairs X = {X1, . . . ,Xn} and Y = {Y1, . . . ,Yn} described in the previous paragraph, let A (resp. B) be the set of indices for the A (resp. B) category. We define XA = 1A ∑a∈A Xa (resp. XB = 1B ∑b∈B Xb) as the average of activation patterns in X with index in A (resp. B). YA (resp. YB) are defined in similar way as the average of activation patterns in Y with index in A (resp. B). Note that the index set is computed from images of X (and not Y ) on both cases. This asymmetry will be useful when designing the permutation scheme. Consider also that we have access to a similarity measure between images that we denote by ρ. For simplicity we will initially suppose that this measure is the Pearson correlation coefficient, although we will see later that this can be generalized to any similarity measure between images. We now have all necessary ingredients to present the test statistic that we propose to distinguish the null hypothesis from 2/17 the alternative. This test statistic has values in [−1,1] and is defined as (cid:16) (cid:124) T (X,Y ) = 1 4 ρ(XA ,YA ) + ρ(XB,YB) − (ρ(XA ,YB) + ρ(XB,YA )) within-condition similarity between-condition similarity (cid:123)(cid:122) (cid:125) (cid:124) (cid:123)(cid:122) (cid:125) (cid:17) . (2) At first, its form might seem strange. Let us give two intuitions on the form of this test statistic: 1. As a difference of similarities. The test statistic can be split as a difference of two terms. The first term is the sum of similarities for images from the same condition (and different modalities), while the second term is a sum of similarities for images of different conditions (and different modalities). Hence, large values of the test statistic are achieved whenever the within-condition similarity is larger than the between-condition similarity, bringing evidence for the existence of a condition-specific activation across modalities. 2. As a singularity test. If we compute all pairwise similarities between the images XA , XB, YA and YB, we obtain 4 scalars that can be arranged in a 2-by-2 matrix as follows: (cid:20)ρ(XA ,YA ) ρ(XA ,YB) (cid:21) ρ(XB,YA ) ρ(XB,YB) . (3) Under the null hypothesis, the samples X and Y are independent and so YA ≈ YB (recall that the indexing was derived from images in X). Whenever YA = YB the matrix above becomes colinear. A standard way to test for colinearity is through its determinant. Computing the determinant of the above equation we obtain our test statistic (modulo the normalizing factor 1 4). Statistical inference. We will estimate the distribution of this test statistic under the null hypothesis from the sample by repeatedly computing the test statistic over a permuted version of the initial sample, a technique often known as permutation or randomization test. For this to be valid, it is necessary to identify the quantities that we wish to permute and verify that under the null hypothesis, all permutations yield the same sample distribution17. Consider the sequence Xπ which results from a random reordering of the activation images in X and the sequence of pairs Z(Xπ ,Y ). Under the null hypothesis, since the probability distribution factorizes over its marginal, the permuted sequence is distributed as P(cid:48) X is the distribution of Xπ. Now, by the iid assumption made previously (which is commonplace in the context of permutation testing), this distribution is invariant to permutations, and so P(cid:48) X = PX and the condition is verified. After computing the permuted test statistic for a large number of random permutations (typically around 10000), the significance of this test, i.e., the probability of observing a test statistic equal or as large as the one obtained, can be computed as X × PY , where P(cid:48) p = number of times{T (Xπ ,Y ) ≥ T (X,Y )} number of permutations . (4) Extensions. This test extends naturally to the setting of group analysis. In this case, the test statistic (2) can be taken as the sum over all subjects of the subject-specific test statistic. Ideally, the same permutation should be used across subjects to obtain each value of the permuted test statistic18. It is theoretically possible to perform a two-tailed test using this test statistic. A large negative value of the test statistic would also bring evidence to reject the null hypothesis of independence. However, since the neuroscientific interpretation of such negative values is not useful for our practical purpose, we will only use the one tailed test in this paper. For simplicity, we have considered ρ the Pearson correlation coefficient as similarity measure, but the method remains valid using any other similarity measures. The Pearson correlation, being a measure of the linear correlation, works best when the effect is (close to) linear, but other more complex similarities can be used such as a (negative) Malahanobis19 or Wasserstein20 distance. Relationship with cross-modal decoding analysis (CMDA). CMDA can be regarded within the same hypothesis testing framework outlined before, but with a different test statistic. In CMDA, the test statistic is the accuracy of a classifier on images from one modality when it was trained on images from the other modality. Since both CMDA and CMPT follow the same permutation test approach to computing significance, both rely implicitly on a label exchangeability assumption behind the data-generating process. As we have seen in the previous subsection, a sufficient condition for this is to assume that the data we observe is sampled iid. Note that this iid assumption is on the pairs from different modalities (Xi,Yi) and also on the experimental paradigm but not on the decoding train/test split, which divides the data by modality and is obviously not iid. This is a much weaker assumption than the distributional assumptions made by 3/17 traditional parametric methods, it is important to keep in mind that permutation tests are not fully assumption-free methods and at the bare minimum require exchangeability of the observations. A practical difference between both approaches is that the cross-modal permutation test is symmetric with respect to modalities while brain decoding is not. That is, CMPT would yield the same p-value regardless of the order in which the different modalities are labeled. This is not true for CMDA, where two possible tests can be performed (train on A and test on B or train on B and test on A), and both can (and typically do) yield different p-values. Datasets Synthetic datasets We construct a synthetic dataset according to a model in which the signal is a superposition of a modality-specific effect (MX ,XY ), a condition-specific effect (Ci) and a Gaussian noise (εi): Xi = α ·Ci + β · MX + εi,Yi = α ·Ci + β · MY + εi where α,β are scalars that regulate the amount of modality-specific and condition-specific signal in the image, respectively. We then generated a total of 20 different images according to this model, considering two different modality-specific signals and two different condition-specific signals, all of them randomly generated from a Gaussian distribution. We generate 3 versions of this dataset, one with 10 voxels, one with 100 and another one with 1000 voxels. fMRI dataset We performed empirical analysis of the data coming from a neurocognitive study of visual attention. fMRI data were collected to investigate object categorization during preparatory activity in a visual search experiment, designed in a similar manner to the one illustrated in21. Participants. 24 participants (8 male, mean age 27.1, st.dev. 4.3 years) were recruited and accessed the research facility. All participantssubjects, before starting the experiment, signed a form confirming their informed consent to participate in the experimental study. After the experiment, they received monetary compensation. Each participant was instructed in advance about undergoing 2 experimental sessions (S1, S2) on two different days. The data on all three modalities in question (perception, imagery and visual search) were acquired in the same session, S1. Out of all 24, 22 completed a significant part (6/22) or the whole (16/22) of S1. During both S1 and S2, participants were also given other tasks, which we do not report here. Out of 16 participants that underwent the whole S1 only nine participants completed 4 runs of both perception/imagery task and visual search task (8 functional runs in total). Other participants failed to reach 4 runs at least in one of the task types. So, for the analysis we are using the data of the 9 participants that have the total of 8 functional runs each. The tasks are explained in the next section. Stimuli. Two distinctive stimulus categories were presented to the participants throughout the tasks: people (whole body image) and cars. Participants were instructed to deal with these categories in three different ways: in perception modality, they had to attend to 8 presentations of 16 seconds long blocks of different instances of the same category (people, here depicted by whole body figures with no face, or cars), interspersed with 16 seconds long fixation periods. Participants were equipped with a two-button box. They were requested to perform a one-back task - i.e., to press a specific button whenever they detected the same image repeated twice in a row. In imagery modality participants were instructed to close their eyes and mentally visualize instances of the category, indicated by the letter cue shown at the beginning of the trial. They had to press the response button whenever they achieved a mental image that was sufficiently detailed, and then they had to switch to mental visualization of another instance of the same category. At the end of the 16 seconds block, an auditory cue told the participants to open their eyes and go on with the experiment. Perception and Imagery blocks were randomly presented within the same functional run. In visual search modality, participants were briefly (450 ms) shown images representing natural scenes (e.g.: crowded places, urban landscapes, etc ...). They were instructed through a visual cue (letter) to look for instances of one of the two categories within the scene. After scene presentation, they had 1.6 s to attend to the presentation of a mask and to give a positive or negative response by pressing a button. Visual search preparatory periods occurring between the presentation of the cue and the presentation of the scene had different lengths: 2, 4, 6, 8 or 10 seconds. Participants did not know in advance neither the lengths of preparatory period, nor their order of presentation throughout the task, which was random. Participants also had to perform a block-designed task, where we alternated presentation of images of intact and scrambled everyday objects, in order to functionally define an object selective region of interest (ROI) localized in the temporal - occipital cortex (Fig. 1). So, for perception and imagery the overall number of trials was 32 per modality (8 trials per 4 runs). Each visual search run consisted of 40 trials, where for each particular type of delay duration there were 8 trials. The total number of visual search trials is 160 (40 per 4 runs), while for each duration there are 32 trials (8 per 4 runs) in the dataset. All experimental procedures had been 4/17 Figure 1 approved by the Ethical Committee of the University of Trento and were carried out in accordance with applicable guidelines and regulations on safety and ethics. Images were acquired with a 4T Bruker (https://www.bruker.com/) scanner. For each participant, we Data acquisition. started both experimental sessions by acquiring a structural scan using a 3D T1-weighted Magnetization Prepared RApid Gradient Echo (MPRAGE) sequence (TR/TE = 2700/4.18 ms, flip angle = 7◦, voxel size = 1 mm isotropic, matrix = 256× 224, 176 sagittal slices). Perception / Imagery and Visual Search tasks were performed while acquiring, respectively, 177 and 195 functional scans with the following parameters: (TR/TE = 2000 / 33 ms, flip angle = 73◦, voxel size = 3x3x3 mm, 1 mm slice spacing, matrix = 64x64, 34 axial slices covering the entire brain), during session 1. Same acquisition parameters were used for 165 scans acquired during Functional Localizer task in session 2. Preprocessing. For data preprocessing FSL tools were used along with in-house built Python code. In all functional runs 5 initial volumes were discarded as dummy volumes. The skull was removed from both functional and structural images to extract the brain. Functional images were subsequently corrected for slice timing and motion artifacts. Transformation of the functional images to standard space was carried out in the following sequence. First structural scans were coregistered to the mean functional scan of each experimental run. Structural-in-functional-space images were then coregistered to standard (MNI) space to finally compute affine parameters. To extract task-related effects from functional localizer data, beta maps for both localizer conditions (Intact vs. Scrambled objects) were computed with linear regression and next fed into contrast analysis (Intact vs. Scrambled). This analysis resulted in a ROI located in bilateral temporal occipital cortex. We selected one cluster including 625 voxels from each hemisphere, ending up with a bilateral ROI with 1250 voxels overall. We applied the ROI mask to functional data coming from Perception / Imagery and Visual Search tasks and we obtained matrices containing time-series of the ROI voxels. For CMPT analysis, 1250x16 matrices of Perception / Imagery data and 1250x40 matrices of Visual Search data were considered per run. Results For the rest of the paper we will refer to the rejection of the null hypothesis with the traditional significance of 0.05 without explicitly mentioning this number. Experiments on synthetic data In Figure 2 we plot the resulting p-value after performing both CMPT and CMDA on the synthetic dataset described in section Methods, for varying magnitudes of the condition-specific effect (α) and different image sizes. In the case of decoding, this p-value was computed as described in27. For α = 0, the dataset has no condition-specific signal and so the test is not expected to produce a statistically significant result. Indeed, the p-value of CMPT is around 0.5. Note that because of the discreteness of the test statistic (test set accuracy), the average p-value need not converge towards 0.5 as α goes to zero in the case for CMDA. As the magnitude of the effect (α) increases, the method that yields a lower p-value has greater statistical power, because it is able to reject the null hypothesis with a greater probability. We can see in the figure, that in general CMDA p-values are higher, which translates into a lower probability of rejecting the null hypothesis under this approach and hence higher Type II error. In Figure 3 (top row) we can see that in the absence of signal (α = 0), the distribution of p-values generated with CMPT(for 6000 repetitions) is relatively flat, showing that the false positive rate (Type I error) for a significance level of β is at the 5/17 Figure 2 Figure 3 6/17 0.00.20.40.60.81.0Magnitude of condition effect ()0.00.10.20.30.40.50.6p-value10 voxels image0.00.20.40.60.81.0Magnitude of condition effect ()0.00.10.20.30.40.50.6100 voxels image0.00.20.40.60.81.0Magnitude of condition effect ()0.00.10.20.30.40.50.61000 voxels imageCMPTDecoding0.05 significance level expected value of β . In the bottom row of that figure, we can see the same experiment for CMDA. In this case because of the discreteness of the test statistic, the distribution is not completely flat. From the simulation results (Figure 2) we see that the average p-values yielded by CMPT are always below those of CMDA. This implies that smaller effects can be detected, and hence, that CMPT has a higher sensitivity than CMDA. Furthermore, this effect is replicated across images with different number of voxels, highlighting the benefits of CMPT in the high-dimensional setting, which is of great practical importance in neuroimaging. Comparison of CMDA and CMPT on fMRI data In this section we assess the agreement or disagreement in detecting shared activation patterns between CMDA and CMPT. The similarity of activation maps for the the discrimination of body vs. car categories was computed for the following pairs of modalities: perception and imagery, imagery and visual search. The visual search modality was investigated more in detail by first considering all durations of preparation periods put together and then analysing separately different delays (2, 4, 6, 8 and 10s). This analysis was meant to emphasize the issue of small sample size typical for neuroscientific data. The comparison between CMDA and CMPT took into account additional elements such as the choice of ROI and the type of encoding of activation maps. The ROI chosen for the analysis was the Object Selective Cortex (OSC) map shown in Fig. 1. We analysed separately the performance of methods for the left part of the ROI, right part of the ROI and the whole ROI. Encoding of the activation maps was of two types: raw BOLD and beta maps. For the raw BOLD encoding, the volumes were selected that corresponded to the peak of the hemodynamic response function (HRF, as rendered by SPM software - www.fil.ion.ucl.ac.uk/spm/doc/) convolved with the boxcar function that represented the experimental manipulation. One volume was selected per trial, and for CMPT the volumes were averaged to produce a single representative volume per subject per condition per modality. For beta encoding, beta maps were calculated trialwise using linear regression; for CMPT the maps were averaged over trials, too, to produce a single beta map per subject per condition per modality. CMDA was performed by training a logistic regression classifier with (cid:96)2 regularization on the trials of one of the modalities. The regularization parameter was selected according to a nested cross-validation scheme (leave-one-run out). The accuracy was then estimated on a test set from another modality. The training and test process was replicated for each subject. Then, the p-value was computed for the group using the permutation scheme described in2. The resulting p-values are reported in Table 1. CMPT group analysis was carried out as described in section Cross-modal permutation test (CMPT). The similarity distance between activation maps that we used is the Pearson correlation measure, both for raw BOLD volume and beta maps encoding. The significance of the proposed test statistics was computed by a permutation scheme with 10.000 iterations to estimate the null distribution. The resulting p-values are reported in Table 1. In Table 1 we report only one result for the comparison between CMDA and CMPT related to raw BOLD volume encoding: the cross-modal analysis between Perception and Imagery. In this case none of the methods detect a meaningful shared activation pattern. Beta maps encoding on the other hand seems to be a more efficient representation. Chen and colleagues22 demonstrate that using beta values is a way to get rid of intrinsic variabilities of BOLD signal throughout the brain and, specifically, within a single area. In our case, this means that beta values are more representative of the effect size than raw BOLD signal changes during task-on periods. For this reason in the presentation of results we only focus on results obtained with data encoded with beta maps. The results in Table 1 confirm our expectations about the presence of common patterns between modalities in Object Selective Cortex. At the same time, CMPT appears to have higher statistical power and sensitivity in revealing these patterns. The results reported in Table 1 illustrate two main scenarios: both CMDA and CMPT show significant p-values or only CMPT. In light of the simulation results we may argue, that since CMDA has a higher false error rate (Figure 2), in case of such disagreements the CMPT result is more reliable. This argument is further supported by the additional empirical evidence that the false positive rate or Type I error is similar for the two tests, limiting the risk of the disagreement being biased by a more optimistic rejection of the null hypothesis. Cross-modal analysis results for perception vs. imagery with beta maps encoding are in agreement between CMDA and CMPT when the two hemispheres are considered individually, namely the left and right OSC respectively. When the analysis is extended to the joint ROI, the number of trials remains constant, while the number of voxels double. In this case the classifier is affected by the higher dimensionality of data, and CMDA does not succeed in rejecting the null hypothesis. The empirical results also support the claim that CMPT is more robust not only in high-dimensional but also in small sample setting. Simulations show that the Type II error of CMPT is below that of CMDA in the small sample regime (Fig. 2). We may find analogous behaviour for the cross-modal analysis of visual search vs. imagery. If we consider the cumulative trials of visual search, irrespective of the delays, CMDA rejects the null hypothesis. When we restrict the cross-modal analysis to single delays of preparation period for visual search, the number of trials drop from 160 to 32. In this case CMDA fails to reject the 7/17 Figure 4 8/17 null hypothesis while CMPT does not. CMPT results are in line with the view that we should expect the presence of shared activity patterns between perceived and imagined object categories. CMPT analysis also confirms that the presence of these patterns can be expected in high-level visual areas processing information about object categories. We are going to further elaborate on this point in the discussion section. On the other hand, CMDA results appear to be affected by the data sample size relative to the high dimensionality of data. Exploratory data analysis with CMPT We ran Searchlight analysis of the whole brain, using CMPT. First, we wanted to show if and how inserting CMPT as the elementary unit within the Searchlight framework could identify voxels that store information related to common activation patterns for two different cognitive modalities. Next, we intended to compare the spatial profiles of the exploratory analysis with the ROI individuated for the confirmatory analysis. To construct group level maps, we referred to the procedure we illustrated in section Cross-modal Permutation Test (CMPT) at a ROI level. Here we consider the spheres centered on each voxel as ROIs: we first compute the "true" statistic and then we proceeded by using permutations. We started by computing single-participants' CMPT - Searchlight maps, where each voxel was considered as the center of a sphere (r= 8), and calculated the T-statistic. Then, we summed up single-participants' T-values and ended up with the true group-level statistic. Next, we created N=10000 permutations of the session labels, and we subsequently constructed 10000 averaged beta maps for each of the conditions based on the permuted labels - that is, two maps per subject per modality per permutation. In this way, we made sure that the data coming from different participants were tested against the same permutations in a uniform way. We then applied CMPT procedure on these permuted maps by first computing an individual T-statistic and then by summing up the group values - that is, we constructed an ad-hoc null distribution. Finally, we simply counted how many times the "true" group statistic was higher than the permuted group statistic, and we transformed the count in a fraction of the total number of permutations, obtaining a p-value. This value was assigned to the voxel in the center of the sphere. The procedure was repeated for each voxel within the gray matter mask. We ended up with a CMPT Searchlight map of p-values coming from a combination of permutation-like tests. In Figures 4 and 5 we present results coming from the exploratory analysis. Fig. 4 showcases the overlap between the whole OSC ROI and informative voxels identified by Searchlight in the occipital-temporal cortex for the cross-modal pair of perception vs. imagery only. In Fig. 5, we put together fragments of maps for the pairs of cognitive processes where our confirmatory analysis yielded significant results, namely perception vs. imagery, visual search (delay 8s) vs. imagery, visual search (all delays) vs. imagery (see Table 1 ). For merely illustrative purposes, the maps were thresholded at the conventional significance level of 0.05 (as we are not aiming at significant cluster identification, no correction for multiple comparisons was carried out). In the left column of Fig. 5 we demonstrate the overlaps between the ROI identified in the course of the group analysis (Fig. 1) and the portions of the map that signal the presence of information about common patterns between two cognitive processes for a single slice (z=-16). The fact that the ROI identified contains a high portion of informative voxels is further illustrated by the histograms in the right column of the same figure. These are histograms of the p-values of the voxels within the ROI. We can see that all three histograms have a skewed shape, signalling the presence of a rather large number of voxels with p-values under 0.05 in the ROI. For comparison, we also ran CMDA Searchlight with the same sphere size (r=8) for the same modality pairs: perception vs. imagery, visual search (all delays) vs. imagery, visual search (delay 8) vs. imagery. The analysis was performed with Matlab 8.5.0, MathWorks, NatickMA, USA using in-house code and Libsvm library (https://www.csie.ntu.edu.tw/ cjlin/libsvm/). Classifier used for producing the maps was an SVM classifier with a linear kernel as implemented in the Libsvm library. For each subject, two Searchlight maps were obtained for each modality pair, one where the classifier was trained on the Imagery data and tested on the other modality data, and one where the assignment of train - test data was reversed. Then, these two maps were averaged as suggested in9 yielding a single map per subject per modality pair. For the group analysis, a one-sample t-test against chance level (50 %) was performed using SPM software. The resulting group maps were thresholded at the significance level of 0.05 and cluster size of 10 voxels. Then we compared the group maps to the OSC ROI selected for the confirmatory analysis. The results are presented in Table 2. It shows percentages of voxels within the OSC ROI that were identified by the CMDA Searchlight as informative about shared patterns between two modalities. The numbers concerning the size of intersection between the Searchlight map and the ROI are given both as an absolute number of voxels within the ROI and in terms of percentages. The problem of asymmetry between classifier results when swapping train and test modalities in a cross-modal setting is well attested for the pair of perception vs. imagery. The accuracies obtained with the classifier trained on the imagery data have been shown to be consistently higher than after training on perception data (28, 33). To minimize the impact of this asymmetry in cross-modal investigations, it was suggested to average the maps resulting from different train-test combinations for a pair of modalities9. However, the authors of the paper showed that the divergence in accuracy numbers obtained with different train-test combinations in their data was insignificant. 9/17 Figure 5 10/17 CMDA Searchlight results in our dataset seem to be rather seriously affected by the issues stemming from the accuracy asymmetries. First, for the pair of perception vs. imagery, CMDA Searchlight trained on imagery data identifies a high number of voxels within the OSC mask, both in right and left OSC. If trained on perception data, the Searchlight finds a much lower number of voxels within the same area, all of them in the right OSC. In the averaged mask, the number of the voxels that survive is nearly 10 times lower than that identified by the Searchlight trained on imagery data (49 against 432). This same kind of asymmetry is even more prominent for the pair of imagery vs. visual search: Searchlight, trained on imagery data, identifies voxels within the OSC ROI, while it does not identify any in the same area if trained with visual search data (neither all delays, nor delay 8). This result makes us pose a question about the extent to which the voxels idenitified belong actually to really shared patterns between modalities, or we should rather talk about voxels in one modality that are informative about the patterns in the other modality. Our overall conclusion about CMDA Searcchlight is that its use might be questionable in cases when notable asymmetry is expected, as is the case with perception vs. imagery. For some modality pairs asymmetry does not seem to be a big issue, as is the case with the data used in9, and the use of CMDA Searchlight could be more justified for these data. Discussion The patterns of brain activity that are shared between the cognitive processes of perception and imagery have been the subject of quite numerous studies. The question investigated was if we can arrive at abstract, top-down object representations28, 29 containing distinguishing features30, 31 that will have common neural substrate both for viewed and imagined object categories30. To test for the presence of shared patterns, many studies used cross-modal decoding - namely, multivariate pattern analysis with SVM classifiers28, 32, 33. Significant cross-modal classification accuracies were taken as the evidence in favour of the presence of shared activity patterns. In some studies, correlation-based analysis was also performed to visualize and estimate similarity between these patterns in terms of distance 28, 32. What emerged from these studies was the view that, indeed, visual imagery activates the same areas that contain information about visually perceived stimuli 30, 34, and shared patterns for stimulus categories in these two processes can be established28, 32, 33, 35 -- 37. The areas where these common representations were found include the ventral temporal pathway, lateral occipital cortex 28, 32, 33, 36 and extrastriate cortex29, 32, 35, 36. The question of shared patterns in early visual areas, such as V1, remains controversial 28, 32. Horikawa31 showed that it depended on the feature type: lower visual features had similar representations for perception and imagery in lower visual areas, while the same was true for higher visual features in higher visual areas. Cichy33 arrived at a similar conclusion about the subdivision of features: although they did not find significant accuracies for decoding object categories in lateral early visual cortex, they could identify shared representations of object locations in these areas. Top-down attention patterns mediate attention biases during perception and affect behavioural performance in attention related tasks. In case of visual attention, these patterns can be revealed in visual search experiments via activity in the category-related object selective areas during preparatory delays38. Several studies attempted at demonstrating the high-level nature of the preparatory patterns through cross-modal analysis, mostly with visual perception as the other modality 21, 38, 39. As object representations in the brain obtained during imagery tasks are thought to be closer to high-level top-down representations of objects in visual cortex 30, 35, the hypothesis naturally suggests itself that we can expect these patterns to show up also during visual search preparatory periods. We tried to shed light on this hypothesis using both CMDA and CMPT on visual imagery and visual search data. Besides, we ran cross-modal analysis separately for preparatory periods of varying length (between 2 and 10 seconds) to get insights into preparatory dynamics. We were expecting that only certain dealys would result significant, conforming different hypotheses about this dynamics. For instance, if only shorter delays (2-4 s) had resulted significant, that could be evidence in favour of transitory and cue-related nature of the preparatory activity in the Object Selective Cortex. If, on the other hand, we had seen significant results in the longer delays, that could reveal the fact that it takes time for the activity to build up. First, we see that both methods confirm expectations about imagery patterns being more high level than perception. None of the methods yielded significant results in the pairs of visual search vs. perception. As for the presence of the shared patterns between visual imagery and visual search, we are faced again with limitations of CMDA as a method: its results can be significant and it can reveal the presence of shared patterns between preparatory periods and imagery, but this type of analysis needs a lot of data. On the other hand, CMPT can reveal shared patterns even with fewer data as is the case with 8 seconds delay. Further study is needed to uncover the temporal dynamics of the preparatory top-down patterns. We hypothesize that delays shorter that 8 s do not allow the preparatory activity to build up, while in case of 10 s the delay it is too long, and the subject might be loosing concentration after a certain period of time. We placed CMPT side by side with other standard data analysis techniques in order to examine whether this approach could be as informative as others. We have shown that in confirmatory, top-down contexts CMPT can yield better results than CMDA. However, it is necessary to mention one limitation of the method. One of the overarching questions in the study of visual imagery is identifying neural representations of categorical features in the form of brain activation maps30, 35. CMPT method cannot provide insights into the location of the discriminative patterns at a ROI level. Despite being a more robust test for cross-modal analysis, CMPT is not appropriate to investigate the shape of shared pattern within a given ROI. In this case, 11/17 CMPT doesn't support a sensitivity analysis at the voxel level needed to compute granular brain maps of activations that are common between modalities. On the other hand, CMDA (at least when linear classifiers are used) contains a vector of weights that can give some clues about the relevance of the input features. However, CMPT combined with Searchlight technique can be a helpful method to locate brain regions that contain information about common patterns between modalities. We took advantage of one strong point of the Searchlight analysis, its "modular" nature: Searchlight might be thought of as a generic framework of data examination that can subsume various analysis techniques as elementary units. Searchlight is widely used in neuroimaging, but it suffers from a number of issues. Conducting Searchlight analysis has several major advantages: first, it can be run on the whole brain, no prior ROI selection is required. Next, it avoids the "curse of dimensionality" of full brain classification, by reducing the number of features used at each point by the classifier. Finally, it has proven to be quite successful in identifying subject specific activation patterns23. The maps produced with Searchlight are of the same nature as the maps obtained with the univariate GLM approach, but they are based on a more fine-grained pattern identification from multiple voxels and better reflect the spatial properties of the BOLD signal (that is, adjacent voxels have similar activation patterns). However, major criticisms of the Searchlight approach regard the use of classification accuracy as the information measure and the t-test as the method to obtain group significances. As is pointed out in23, SVM classifiers can correctly classify even with a few number of highly informative voxels and when weakly informative voxels are numerous enough. Both of these behaviours can cause distortions in a map: in the first case, all searchlights overlapping with one of a few informative voxels will be significant. In this way, the number of informative voxels is overestimated. In the second case, the cause of distortions is "discontinuous information detection": groups of weakly informative voxels will be missed out if their size is below a certain threshold, but can be judged significant if you just add a single voxel. That leads to underestimation of the number of significant clusters just because the number of weakly significant voxels does not reach a certain mass. Efficiency of using classifiers with Searchlight depends strongly on the classifier parameters and sphere size23, 24. In24, the point is raised against interpretability of classifier accuracy with neuroscientific data: unlike distance measures, its value depends on the properties of the dataset (amount of training data and what kind of data is used as test data) and not only on the presence of a particular effect in the data. Besides, the authors point out that capturing interactions of several factors in a factorial experimental design cannot be cast as a classification task. So, addressing these methodological issues for Searchlight can significantly improve this valuable tool and make its result more scientifically rigorous. Classification accuracy is not the only way to represent information content. In the original paper by Kriegeskorte 25 the metrics used was Mahalanobis distance between the distributions corresponding to stimulus categories. In 24 the authors build on the probabilistic model of the data proposing a cross-validated multivariate ANOVA (MANOVA) as the informational content measure. In19 three various measures - classification accuracy, Euclidean/Mahalanobis distance, and Pearson correlation distance - are compared for reliability in the context of Searchlight analysis. In this paper, it was shown that "continuous crossvalidated distance estimators" such as Euclidean/Mahalanobis distance or Pearson correlation should be preferred for Searchlight because they are more interpretable from the neuroscientific viewpoint. Another bunch of critical remarks concerns the use of t-tests for assessing significance at the group level. Certain properties of neuroscientific data make the use of t-tests questionable for this purpose, "particularly, the low number of observations and the non-gaussianity of the probability distribution of accuracy. As a consequence, several assumptions of the t-statistic are not met, rendering the procedure invalid from a theoretical point of view" 26. However, it is not the only option here. In26 a non-parametric test for group significance and cluster inference was proposed based on permutations and bootstrapping procedure. Nastase and colleagues9 also opt for permutation tests in Searchlight context. Methodologically, using CMPT in conjunction with the Searchlight technique for cross-modal pattern analysis has several advantages over the common Searchlight procedure because it does not rely neither on classification accuracy nor on the t-tests and hence avoids the common methodological pitfalls. At the same time, we are following the suggestions in the literature that are considered more appropriate for the Searchlight. First, the test statistic proposed in equation 2 that is used as the measure of information contained at each voxel is based on Pearson correlation and is interpretable in terms of similarity. Second, group significance is tested non-parametrically with permutation tests that do not make assumptions about the shape of the data distribution. We found that CMPT integrated into Searchlight has proven effective also to explore and, potentially, confirm what we observed using top-down, ROI-based analysis, which suggests both robustness and efficiency of the CMPT Searchlight in fMRI data analysis. However, it is important to note that confirmatory and exploratory analyses report different p-values. While it is possible to qualitatively compare the outcomes of these two analyses, plainly putting their p-values side by side might be misleading. CMPT-ROI p-values come from an extended, functionally well-defined area including 625 or 1250 voxels. CMPT-SL analysis spans over the whole brain sphere by sphere, extracting results from spheres including about 200 voxels each. This means that p-values coming from confirmatory and exploratory analysis should not be compared on a purely quantitative level. The question of shared patterns between various cognitive modalities is relevant not only for object categorization in visual processing. It is fundamental in the study of interactions between top-down and bottom-up processing streams in the human 12/17 brain in general. Further directions of study could include using the CMPT method and the CMPT Searchlight technique with a wider number of other cognitive modalities, such as auditory or linguistic40 -- 42. Besides, we could investigate other areas that can share representations with imagery - for instance, working memory areas35. Finally, the CMPT method could be tried with other types of neuroimaging data - as, for example, EEG motor imagery data for Brain-Computer interfaces 43 or MEG data 44, 45. Acknowledgements The research was partially funded by the Autonomous Province of Trento, Call "Grandi Progetti 2012", project "Characterizing and improving brain mechanisms of attention - ATTEND" and by Centro Internazionale per la Ricerca Matematica (CIRM). The authors gratefully acknowledge the contribution of Marius Peelen to the disucssions of experiment design and data analysis and to Valentina Borghesani for valuable feedback on the manuscript. Author contributions statement E.K., V.I. and P.A participated in designing the experiment. E.K. and V.I. implemented the design, conducted the experiment, acquired the data and preprocessed the data. All authors analysed the data. E.K., F.P. and E.O. implemented data analysis pipelines. F.P. designed, implemented and conducted experiments on simulated data. E.K., F.P., V.I. and P.A. wrote the manuscript. Additional information The authors declare no competing interests. Data and code availability. Processed data (ROI beta-maps, BOLD volumes and Searchlight maps) are available from a public repository on Github https://github.com/elena-kalinina/Code_CMPT_paper. The code used to generate and analyze synthetic data is available from the same repository along with data analysis code and Searchlight implementation. Sharing the whole dataset involves the consent of third-parties who participated in the experiment design. If the consensus is reached, the data could be made publicly available. References 1. Knops, A., Thirion, B., Hubbard, E. M., Michel, V. & Dehaene, S. Recruitment of an area involved in eye movements during mental arithmetic. Sci. 324, 1583 -- 1585 (2009). 2. Etzel, J. A., Gazzola, V. & Keysers, C. Testing simulation theory with cross-modal multivariate classification of fMRI data. PLoS ONE 3(11), URL http://dx.doi.org/10.1371/journal.pone.0003690. DOI 10.1371/jour- nal.pone.0003690. (2008). 3. Shinkareva, S. V., Malave, V. L., Mason, R. A., Mitchell, T. M. & Just, M. A. Commonality of neural representations of words and pictures. NeuroImage 54, 2418 -- 2425 (2011). 4. Haynes, J.-D. & Rees, G. Decoding mental states from brain activity in humans. Nat. Rev. Neurosci. 7, 523 -- 534 (2006). 5. Kaplan, J. T., Man, K. & Greening, S. G. Multivariate cross-classification: applying machine learning techniques to charac- terize abstraction in neural representations. Front. Hum. Neurosci. 9, 151 URL http://journal.frontiersin. org/article/10.3389/fnhum.2015.00151. DOI 10.3389/fnhum.2015.00151. (2015). 6. Majerus, S. et al. Cross-modal decoding of neural patterns associated with working memory: evidence for attention-based accounts of working memory. Cereb. cortex 26, 166 -- 179 (2016). 7. Vetter, P., Smith, F. W. & Muckli, L. Decoding sound and imagery content in early visual cortex. Curr. Biol. 24, 1256 -- 1262 (2014). 8. Kaiser, D., Azzalini, D. C. & Peelen, M. V. Shape-independent object category responses revealed by MEG and fMRI decoding. J. Neurophysiol. 115(4), 2246 -- 50 (2016). 9. Nastase, S. A., Halchenko, Y. O., Davis, B. & Hasson, U. Cross-modal searchlight classification: methodological challenges and recommended solutions. In Pattern Recognition in NeuroImaging (PRNI), 2016 International Workshop on, 1 -- 4. IEEE. URL https://ieeexplore.ieee.org/document/7552355. DOI 10.1109/PRNI.2016.7552355. (2016). 13/17 10. Fisher, R. A. The Design of Experiments. Oliver Boyd, Edinb. (1935). 11. Pitman, E. J. Significance tests which may be applied to samples from any populations. Suppl. to J. Royal Stat. Soc. 4, 119 -- 130 (1937). 12. Lehmann, E. L. & Stein, C. On the theory of some non-parametric hypotheses. The Annals Math. Stat. 20(1), 28 -- 45 (1949). 13. Nichols, T. E. & Holmes, A. P. Nonparametric permutation tests for functional neuroimaging: a primer with examples. Hum. brain mapping 15, 1 -- 25 (2002). 14. Eklund, A., Nichols, T. E. & Knutsson, H. Cluster failure: why fmri inferences for spatial extent have inflated false-positive rates. Proc. Natl. Acad. Sci. 113(28), 7900 -- 7905 (2016). 15. Woolrich, M. W., Beckmann, C. F., Nichols, T. E. & Smith, S. M. Statistical analysis of fMRI data. In fMRI Techniques and Protocols 41, 183 -- 239 Springer-Verlag, New York. (2009). 16. Winkler, A. M., Ridgway, G. R., Douaud, G., Nichols, T. E. & Smith, S. M. Faster permutation inference in brain imaging. NeuroImage 141, 502 -- 516 (2016). 17. Lehmann, E. L. & Romano, J. P. Testing Statistical Hypotheses. Springer-Verlag New York. (2005). 18. Etzel, J. A. MVPA permutation schemes: permutation testing for the group level. In Pattern Recognition in NeuroImaging (PRNI), 2015 International Workshop on, 65 -- 68. IEEE. URL https://ieeexplore.ieee.org/document/ 7270849. DOI 10.1109/prni.2015.29. (2015) 19. Walther, A. et al. Reliability of dissimilarity measures for multi-voxel pattern analysis. NeuroImage 137, 188 -- 200 (2016). 20. Gramfort, A., Peyr´e, G. & Cuturi, M. Fast optimal transport averaging of neuroimaging data. Preprint at URL http: //arxiv.org/abs/1503.08596. 1503.08596. (2015). 21. Peelen, M. V., Fei-Fei, L. & Kastner, S. Neural mechanisms of rapid natural scene categorization in human visual cortex. Nat. 460, 94 -- 97 (2009). 22. Chen, X., Pereira, F., Lee, W., Strother, S. & Mitchell, T. Exploring predictive and reproducible modeling with the single-subject FIAC dataset. Hum. brain mapping 27, 452 -- 461 (2006). 23. Etzel, J. A., Zacks, J. M. & Braver, T. S. Searchlight analysis: promise, pitfalls, and potential. NeuroImage 78, 261 -- 269 (2013). 24. Allefeld, C. & Haynes, J.-D. Searchlight-based multi-voxel pattern analysis of fMRI by cross-validated MANOVA. NeuroImage 89, 345 -- 357 (2014). 25. Kriegeskorte, N., Goebel, R. & Bandettini, P. Information-based functional brain mapping. Proc. Natl. Acad. Sci. United States Am. 103, 3863 -- 3868 (2006). 26. Stelzer, J., Chen, Y. & Turner, R. Statistical inference and multiple testing correction in classification-based multi-voxel pattern analysis (MVPA): Random permutations and cluster size control. NeuroImage 65, 69 -- 82 (2013). 27. Ojala, Markus, Garriga, Gemma C, Permutation tests for studying classifier performance. J. Mach. Learn. Res. 11, 1833 -- 1863 (2010) 28. Reddy, L., Tsuchiya, N. & Serre, T. Reading the mind's eye: decoding category information during mental imagery. NeuroImage 50, 818 -- 825 (2010). 29. Ishai, A. Seeing faces and objects with the "mind's eye". Arch. italiennes de biologie 148, 1 -- 9 (2010). 30. Roldan, S. M. Object recognition in mental representations: directions for exploring diagnostic features through visual mental imagery. Front. Psychol. 8 URL http://dx.doi.org/10.3389/fpsyg.2017.00833. DOI 10.3389/fp- syg.2017.00833. (2017). 31. Horikawa, T. & Kamitani, Y. Generic decoding of seen and imagined objects using hierarchical visual features. Nat. Commun. 8 URL http://dx.doi.org/10.1038/ncomms15037. DOI 10.1038/ncomms15037. (2017). 32. Lee, S.-H., Kravitz, D. J. & Baker, C. I. Disentangling visual imagery and perception of real-world objects. NeuroImage 59, 4064 -- 4073 (2012). 33. Cichy, R. M., Heinzle, J. & Haynes, J.-D. Imagery and perception share cortical representations of content and location. Cereb. Cortex 22, 372 -- 380 (2012). 34. Farah, M. J. The neural basis of mental imagery. Trends neurosciences 12, 395 -- 399 (1989). 14/17 35. Pearson, J., Naselaris, T., Holmes, E. A. & Kosslyn, S. M. Mental imagery: functional mechanisms and clinical applications. Trends cognitive sciences 19, 590 -- 602 (2015). 36. Stokes, M., Thompson, R., Cusack, R. & Duncan, J. Top-down activation of shape-specific population codes in visual cortex during mental imagery. The J. neuroscience : official journal Soc. for Neurosci. 29, 1565 -- 1572 (2009). 37. Anderson, A. J. J., Bruni, E., Lopopolo, A., Poesio, M. & Baroni, M. Reading visually embodied meaning from the brain: Visually grounded computational models decode visual-object mental imagery induced by written text. NeuroImage 120, 309 -- 322 (2015). 38. Stokes, M., Thompson, R., Nobre, A. C. & Duncan, J. Shape-specific preparatory activity mediates attention to targets in human visual cortex. Proc. Natl. Acad. Sci. 106, 19569 -- 19574 (2009). 39. Peelen, M. V. & Kastner, S. A neural basis for real-world visual search in human occipitotemporal cortex. Proc. Natl. Acad. Sci. United States Am. 108, 12125 -- 12130 (2011). 40. Simanova, I., Hagoort, P., Oostenveld, R. & van Gerven, M. A. Modality-independent decoding of semantic information from the human brain. Cereb. cortex 24, 426 -- 434 (2014). 41. Simanova, I., Francken, J. C., de Lange, F. P. & Bekkering, H. Linguistic priors shape categorical perception. Lang. Cogn. Neurosci. 31, 159 -- 165 (2016). 42. Borghesani, V., Pedregosa, F., Eger, E., Buiatti, M. & Piazza, M. A perceptual-to-conceptual gradient of word coding along the ventral path. In Pattern Recognition in Neuroimaging, 2014 International Workshop on, 1 -- 4. IEEE. URL https://ieeexplore.ieee.org/document/6858512. DOI 10.1109/prni.2014.6858512. (2014) 43. Choi, K. Electroencephalography (EEG)-based neurofeedback training for brain-computer interface (BCI). Exp. brain research 231, 351 -- 365 (2013). 44. Dikker, S. & Pylkkanen, L. Predicting language: MEG evidence for lexical preactivation. Brain language 127, 55 -- 64 (2013). 45. Hirschfeld, G., Zwitserlood, P. & Dobel, C. Effects of language comprehension on visual processing - MEG dissociates early perceptual and late N400 effects. Brain language 116, 91 -- 96 (2011). Modalities P vs I P vs I P vs I P vs I V vs I (delay 2 s) V vs I (delay 4 s) V vs I (delay 6 s) V vs I (delay 8 s) V vs I (delay 10 s) V vs I (all delays) ROI L+R L+R L R L+R L+R L+R L+R L+R L+R Encoding CMDA CMPT 0.75 0.001 0.001 0.008 0.589 0.141 0.303 0.004 0.461 0.001 0.12 0.08 0.001 0.005 0.272 0.387 0.165 0.659 0.249 0.012 Bold Beta Beta Beta Beta Beta Beta Beta Beta Beta Table 1. Comparison of cross-modal decoding analysis (CMDA) and cross-modal permutation test (CMPT). The values refer to p-value for the discrimination task of body vs. car in each pair of the cognitive modalities listed in the extreme left column, where P stands for perception, I for imagery (I) and V for visual search. 15/17 Modalities P vs I P vs I P vs I V vs I V vs I V vs I P vs I P vs I P vs I V vs I V vs I V vs I P vs I P vs I P vs I V vs I V vs I V vs I V vs I (delay 8) V vs I (delay 8) V vs I (delay 8) V vs I (delay 8) V vs I (delay 8) V vs I (delay 8) V vs I (delay 8) V vs I (delay 8) V vs I (delay 8) mean mean mean single P I single I P single V I single I V single I V single V I ROI Map Type L L L L L L L L L R R R R R R R R L single V I single I V single I V single V I single P I single I P mean mean mean mean mean single P I single I P L+R L+R L+R L+R L+R single V I L+R single I V L+R L+R single I V L+R single V I mean Intersection N Vox Intersection % 4 0 239 5 0 12 0 29 0 45 13 193 0 0 1 0 10 0 49 13 432 5 0 13 0 39 0 0.64 0 38.24 0.8 0 1.92 0 4.64 0 7.2 2.08 30.88 0 0 0.16 0 1.6 0 3.92 1.04 34.56 0.4 0 1.04 3.12 0 0 . Table 2. Percentage of voxels within the OSC ROI identified by the CMDA Searchlight as informative about shared patterns between two modalities. The cognitive modalities are listed in the extreme left column. P stands for perception, I for imagery and V for visual search. In the second column, it is shown which part of the OSC mask was used for calculations: left (L), right (R) or the whole mask (L+R). Number of voxels in L+R ROI is equal to 1250; each half of it (both L and R) has 650 voxels. The third column shows which type of map was used for calculations: an averaged map (denoted as mean), or a single training - test pair map (denoted as single). In the names of the single maps, first comes the training modality, and second the test modality. In the last two columns you can find the numbers concerning the size of intersection between the map in column 3 and the ROI in column 2. In column 4, the intersection size is given as an absolute number of voxels within the ROI. In the last, extreme right column, it is given in terms of percentages. Figure 1. Object Selective Cortex (OSC) group map in temporal-occipital cortex delineated based on the functional localizer (intact vs. scrambled objects). Its right (R) hemisphere part is shown in red, left (L) in blue. X, Y and Z locate the coordinates of the slices. Figure 2. Statistical power as a function of the effect size. We compare the obtained p-value (y) for different magnitudes of the condition-specific effect (x) and different image sizes (10, 100 and 1000 voxels). Shaded areas correspond to the standard deviation. Across these different scenarios, CMPT yields a higher significance than the decoding-based approach. Figure 3. The distribution of p-values generated with CMPT (top) and CMDA (bottom) in the absense of condition-specific effect after 6000 repetitions of the experiment. p-values are on the x axis, while the frequencies are shown on the y axis. In case of CMPT (top) the distribution converges to the uniform. The false positive rate (Type I error) for a significance level of β is at the expected value of β . For CMDA (bottom), the distribution is not fully flat due to the discreteness of the scoring rule. Figure 4. Overlap between the whole OSC ROI and informative voxels identified by Searchlight in the occipital-temporal cortex for the cross-modal pair of perception vs. imagery. The ROI (the same as in Figure fig:roi) is shown in blue. P-values in the Searchlight map are colorcoded. The maps were thresholded at the conventional significance level of 0.05, and the color code shows values between 0.95 (corresponding to the p-value of 0.05) in red and 1 (corresponding to the p-value of 0.00) in yellow. The leftmost column of the figure shows sagitarial (top) and coronal (bottom) slices (at x= 42 and y = -74 correspondingly) where the rectangle shaded in blue delimits the axial slices further presented in columns 2 - 4. These are six 16/17 slices of the axial plane taken at steps of 8 between coordinates z = -24 and z = 16. We can see that in all these six slices, there are overlaps between the OSC ROI and areas where Searchlight identified voxels informative about shared activity patterns between perception and imagery. Figure 5. CMPT - Searchlight analysis results. The rows represent results coming from the pairs of cognitive processes where our confirmatory analysis yielded significant p-values (Table 1): perception vs. imagery (top), visual search (delay of 8 s) vs. imagery (middle), visual search (all delays) vs. imagery (bottom). In each row, the left column shows the overlap between the Object Selective Cortex ROI previously defined for confirmatory analysis (Fig. 1) and the Searchlight maps obtained for this pair. In the right column, there is a histogram of p-values within the OSC ROI coming from the Searchlight map. The left column showcases axial slices of Searchlight maps for the corresponding pairs, focusing on posterior region. Slices are taken at z= -16, maps were thresholded at the significance level of 0.05, and the color code shows values between 0.95 (corresponding to the p-value of 0.05) in red and 1 (corresponding to the p-value of 0.00) in yellow. The OSC ROI is presented in green. Overlaps between the green area and colorcoded maps can be observed for all three pairs. The right column further illustrates the fact that the confirmatory OSC ROI identified contains a high portion of informative voxels. In each histogram, the x axis represents the p-values, while their frequencies (in % from the overall number of voxels in the area) are ordered along the y-axis. All three histograms are skewed to the right, signalling the presence of a rather large number of voxels with p-values under 0.05 in the ROI. 17/17
1307.5728
2
1307
2013-11-07T10:39:01
How adaptation currents change threshold, gain and variability of neuronal spiking
[ "q-bio.NC" ]
Many types of neurons exhibit spike rate adaptation, mediated by intrinsic slow $\mathrm{K}^+$-currents, which effectively inhibit neuronal responses. How these adaptation currents change the relationship between in-vivo like fluctuating synaptic input, spike rate output and the spike train statistics, however, is not well understood. In this computational study we show that an adaptation current which primarily depends on the subthreshold membrane voltage changes the neuronal input-output relationship (I-O curve) subtractively, thereby increasing the response threshold. A spike-dependent adaptation current alters the I-O curve divisively, thus reducing the response gain. Both types of adaptation currents naturally increase the mean inter-spike interval (ISI), but they can affect ISI variability in opposite ways. A subthreshold current always causes an increase of variability while a spike-triggered current decreases high variability caused by fluctuation-dominated inputs and increases low variability when the average input is large. The effects on I-O curves match those caused by synaptic inhibition in networks with asynchronous irregular activity, for which we find subtractive and divisive changes caused by external and recurrent inhibition, respectively. Synaptic inhibition, however, always increases the ISI variability. We analytically derive expressions for the I-O curve and ISI variability, which demonstrate the robustness of our results. Furthermore, we show how the biophysical parameters of slow $\mathrm{K}^+$-conductances contribute to the two different types of adaptation currents and find that $\mathrm{Ca}^{2+}$-activated $\mathrm{K}^+$-currents are effectively captured by a simple spike-dependent description, while muscarine-sensitive or $\mathrm{Na}^+$-activated $\mathrm{K}^+$-currents show a dominant subthreshold component.
q-bio.NC
q-bio
How adaptation currents change threshold, gain and variability of neuronal spiking Josef Ladenbauer1,2∗, Moritz Augustin1,2, Klaus Obermayer1,2 1 Neural Information Processing Group, Technische Universitat Berlin, Berlin, Germany 2 Bernstein Center for Computational Neuroscience Berlin, Berlin, Germany ∗Correspondence: Josef Ladenbauer Technische Universitat Berlin Department of Software Engineering and Theoretical Computer Science Neural Information Processing Marchstr. 23, MAR 5049 10587 Berlin, Germany E-mail: [email protected] Abstract Many types of neurons exhibit spike rate adaptation, me- diated by intrinsic slow K+-currents, which effectively in- hibit neuronal responses. How these adaptation currents change the relationship between in-vivo like fluctuating synaptic input, spike rate output and the spike train statis- tics, however, is not well understood. In this computa- tional study we show that an adaptation current which primarily depends on the subthreshold membrane volt- age changes the neuronal input-output relationship (I- O curve) subtractively, thereby increasing the response threshold, and decreases its slope (response gain) for low spike rates. A spike-dependent adaptation current alters the I-O curve divisively, thus reducing the response gain. Both types of adaptation currents naturally increase the mean inter-spike interval (ISI), but they can affect ISI vari- ability in opposite ways. A subthreshold current always causes an increase of variability while a spike-triggered current decreases high variability caused by fluctuation- dominated inputs and increases low variability when the average input is large. The effects on I-O curves match those caused by synaptic inhibition in networks with asyn- chronous irregular activity, for which we find subtractive and divisive changes caused by external and recurrent inhi- bition, respectively. Synaptic inhibition, however, always increases the ISI variability. We analytically derive expres- sions for the I-O curve and ISI variability, which demon- strate the robustness of our results. Furthermore, we show how the biophysical parameters of slow K+-conductances contribute to the two different types of adaptation currents and find that Ca2+-activated K+-currents are effectively captured by a simple spike-dependent description, while muscarine-sensitive or Na+-activated K+-currents show a dominant subthreshold component. Introduction Adaptation is a widespread phenomenon in nervous sys- tems, providing flexibility to function under varying ex- ternal conditions. At the single neuron level this can be observed as spike rate adaptation, a gradual decrease in spiking activity following a sudden increase in stimulus intensity. This type of intrinsic inhibition, in contrast to the one caused by synaptic interaction, is typically medi- ated by slowly decaying somatic K+-currents which accu- mulate when the membrane voltage increases. A number of slow K+-currents with different activation characteris- tics have been identified. Muscarine-sensitive (Brown and Adams 1980; Adams et al. 1982) or Na+-dependent K+- channels activate at subthreshold voltage values (Schwindt et al. 1989; Kim and McCormick 1998), whereas Ca2+- dependent K+-channels activate at higher, suprathreshold values (Brown and Griffith 1983; Madison and Nicoll 1984; Schwindt et al. 1992). Such adaptation currents, for example, mediate fre- quency selectivity of neurons (Fuhrmann et al. 2002; Benda et al. 2005; Ellis et al. 2007), where the preferred frequency depends on the current activation type (Dee- myad et al. 2012). They promote network synchroniza- tion (Sanchez-Vives and McCormick 2000; Augustin et al. 2013; Ladenbauer et al. 2013) and are likely involved in the attentional modulation of neuronal response proper- ties by acetylcholine (Herrero et al. 2008; Soma et al. 2012; McCormick 1992). It has been hypothesized that these complex effects are produced by changing the relationship between synaptic input and spike rate output (I-O curve) (Deemyad et al. 2012; Benda and Herz 2003; Soma et al. 2012; Reynolds and Heeger 2009). For example, changing the I-O curve of a neuron subtractively sharpens stimu- lus selectivity, whereas a divisive change downscales the neuronal response but preserves selectivity (see (Wilson et al. 2012) in the context of synaptic inhibition). It was also suggested, that adaptation currents affect the neu- ral code via their effect on the inter-spike interval (ISI) statistics (Prescott and Sejnowski 2008). So far, effects of adaptation currents on I-O curves have been studied con- sidering constant current inputs disregarding input fluc- tuations (Prescott and Sejnowski 2008; Deemyad et al. 2012) and it has remained unclear how different types of adaptation currents affect ISI variability. Therefore, in this contribution we systematically examine how voltage- dependent subthreshold and spike-dependent adaptation currents change neuronal I-O curves as well as the ISI dis- tribution for typical in-vivo like input statistics, and how the biophysical parameters of slow K+-conductances con- tribute to the two types of adaptation currents. We address these questions by studying spike rates and ISI distributions of model neurons with subthreshold and spike-triggered adaptation currents, subject to fluctuat- ing in-vivo like inputs, and we compare the results to those induced by synaptic inhibition. Specifically, we use the adaptive exponential integrate-and-fire (aEIF) neuron model (Brette and Gerstner 2005), which has been shown to perform well in predicting the subthreshold properties (Badel et al. 2008) and spiking activity (Jolivet et al. 2008; Pospischil et al. 2011) of cortical neurons. To an- alytically demonstrate the changes of I-O curves and ISI variability we derive explicit expressions for these proper- ties based on the simpler perfect integrate-and-fire neuron model (see, e.g., (Gerstein and Mandelbrot 1964)) with adaptation (aPIF). Finally, using a detailed conductance- based neuron model we quantify the subthreshold and spike-triggered components of various slow K+-currents and compare the effects of specific K+-channels on the 3 1 0 2 v o N 7 ] . C N o i b - q [ 2 v 8 2 7 5 . 7 0 3 1 : v i X r a 1 I-O curve and ISI variability. Materials and Methods aEIF neuron with noisy input current We consider an aEIF model neuron receiving synaptic in- put currents. The subthreshold dynamics of the mem- brane voltage V is given by C dV dt = Iion(V ) + Isyn(t), (1) where the capacitive current through the membrane with capacitance C equals the sum of ionic currents Iion and the synaptic current Isyn. Three ionic currents are taken into account, Iion(V ) := −gL(V −EL)+gL ∆T exp −w. (2) (cid:18) V − VT (cid:19) ∆T The first term on the right-hand side describes the leak current with conductance gL and reversal potential EL. The exponential term with threshold slope factor ∆T and effective threshold voltage VT approximates the fast Na+- current at spike initiation, assuming instantaneous activa- tion of Na+-channels (Fourcaud-Trocm´e et al. 2003). w is the adaptation current which reflects a slow K+-current. It evolves according to τw dw dt = a(V − Ew) − w, (3) with adaptation time constant τw. Its strength depends on the subthreshold membrane voltage via conductance a. Ew denotes its reversal potential. When V increases be- yond VT, a spike is generated due to the exponential term in eq. (2). The downswing of the spike is not explicitly modelled, instead, when V reaches a value Vs ≥ VT, the membrane voltage is reset to a lower value Vr. At the same time, the adaptation current w is incremented by a value of b, implementing the mechanism of spike-triggered adap- tation. Immediately after the reset, V and w are clamped for a refractory period Tref , and subsequently governed again by eqs. (1) -- (3). The aEIF model range of neuronal and Brette al. 2008). We selected the following parameter val- C = 1 µF/cm2, ues gL = 0.05 mS/cm2, ∆T = 1.5 mV, Ew = −80 mV, VT = −50 mV, Vs = −40 mV, Vr = −70 mV and Tref = 1.5 ms (Badel et al. 2008; Destexhe 2009; Wang et al. 2003). The adap- tation parameters a and b were varied within reasonable ranges, a ∈ [0, 0.06] mS/cm2, b ∈ [0, 0.3] µA/cm2. neurons: EL = −65 mV, τw = 200 ms, subthreshold dynamics patterns can reproduce a wide (Touboul (Naud et to model cortical 2008) and spike The synaptic input consists of a mean µ(t) and a fluc- tuating part given by a Gaussian white noise process η(t) with δ-autocorrelation and standard deviation σ(t), Isyn(t) = C [µ(t) + σ(t)η(t)] . (4) Eq. (4) describes the total synaptic current received by KE excitatory and KI inhibitory neurons, which produce instantaneous postsynaptic potentials (PSPs) JE > 0 and JI < 0, respectively. For synaptic events (i.e. presynaptic 2 spike times) generated by independent Poisson processes with rates rE (t) and rI(t), the infinitesimal moments µ(t) and σ(t) are expressed as µ(t) = JE KE rE (t) + JIKIrI(t), σ(t)2 = J 2E KE rE (t) + J 2IKIrI(t), (5) (6) assuming large numbers KE , KI and small magnitudes of JE , JI (Tuckwell 1988; Renart et al. 2004; Destexhe and Rudolph-Lilith 2012). This diffusion approximation well describes the activity in many cortical areas (Shadlen and Newsome 1998; Destexhe et al. 2003; Compte et al. 2003; Maimon and Assad 2009). The parameter values were JE = 0.15 mV, JI = −0.45 mV, KE = 2000, KI = 500 and rE , rI were varied in [0, 50] Hz. In addition, we directly varied µ and σ over a wide range of biologically plausible values. Membrane voltage distribution and spike rate In the following we describe how we obtain the distribu- tion of the membrane voltage p(V, t) and the instantaneous spike rate r(t) of a single neuron at time t for a large num- ber N of independent trials. Note that by trial we refer to a solution trajectory of the system of stochastic differential equations eq. (1) -- (4) for a realization of η(t). ¯w(t) 1/N(cid:80)N First, to reduce computational demands and enable further analysis, we replace the adaptation current w in eqs. (2) -- (3) by its average over trials, := i=1 wi(t), where i is the trial index (Gigante et al. 2007a). Neglecting the variance of w across trials is valid under the assumption that the dynamics of the adapta- tion current is substantially slower than that of the mem- brane voltage, which is supported by empirical observa- tions (Brown and Adams 1980; Sanchez-Vives and Mc- Cormick 2000; Sanchez-Vives et al. 2000; Stocker 2004). The instantaneous spike rate at time t can be estimated by the average number of spikes in a small interval [t, t + ∆t], (cid:90) t+∆t N(cid:88) (cid:88) i=1 t k r∆t(t) := 1 N ∆t δ(s − tk i )ds, (7) where δ is the delta function and tk i denotes the k-th spike time in trial i. In the limit N → ∞, ∆t → 0, the prob- ability density p(V, t) obeys the Fokker-Planck equation (Risken 1996; Tuckwell 1988; Renart et al. 2004), p(V, t) + q(V, t) = 0, (8) with probability flux q(V, t) given by ∂ ∂t C (cid:18) Iion(V ; ¯w) ∂ ∂V (cid:19) q(V, t) := + µ(t) p(V, t) − σ(t)2 2 ∂ ∂V p(V, t). (9) Iion(V ; ¯w) denotes the sum of ionic currents (cf. eq. (2)) where w is replaced by the average adaptation current ¯w which evolves according to τw d ¯w dt = a((cid:104)V (cid:105)p(V,t) − Ew) − ¯w + τw b r(t). (10) (cid:104)·(cid:105)p indicates the average with respect to the probability density p (Brunel et al. 2003; Gigante et al. 2007b). To account for the reset of the membrane voltage, the prob- ability flux at Vs is re-injected at Vr after the refractory period has passed, i.e., q(V, t) − lim V (cid:37)Vr lim V (cid:38)Vr q(V, t) = q(Vs, t − Tref ). (11) The boundary conditions for this system are reflecting for V → −∞ and absorbing for V = Vs, V →−∞ q(V, t) = 0, lim p(Vs, t) = 0, (12) and the (instantaneous) spike rate is obtained by the prob- ability flux at Vs, r(t) = q(Vs, t). (13) Note that p(V, t) only reflects the proportion of trials where the neuron is not refractory at time t, given by P (t) =(cid:82) Vs−∞ p(v, t)dv (< 1 for Tref > 0 and r(t) > 0). The eq. (10) is calculated as (cid:104)V (cid:105)p(V,t) =(cid:82) Vs−∞ vp(v, t)dv/P (t). total probability density that the membrane voltage is V at time t is given by p(V, t) + pref (V, t), with refractory density pref (V, t) = (1 − P (t)) δ(V − Vr). Since p(V, t) does not integrate to unity in general, the average in The dynamics of the average adaptation current ¯w(t) reflecting the non-refractory proportion of trials is well captured by eq. (10) as long as Tref is small compared to τw. In this (physiologically plausible) case ¯w(t) can be considered equal to the average adaptation current over the refractory proportion of trials. Steady-state spike rate We consider the membrane voltage distribution of an aEIF neuron with noisy synaptic input, described by the equa- tions (8) -- (13), has reached its steady-state p∞. p∞ obeys ∂p∞(V )/∂t = 0 or equivalently, q∞(V ) = 0, (14) ∂ ∂V (cid:18) Iion(V ; ¯w) C (cid:19) with steady-state probability flux q∞ given by q∞(V ) = + µ p∞(V )− σ2 2 ∂ ∂V p∞(V ), (15) subject to the reset condition, q∞(V ) − lim V (cid:37)Vr lim V (cid:38)Vr and the boundary conditions, q∞(V ) = q∞(Vs), (16) V →−∞ q∞(V ) = 0 lim p∞(Vs) = 0 (17) The steady-state spike rate is given by r∞ = q∞(Vs) and the steady-state mean adaptation current reads ¯w∞ = a((cid:104)V (cid:105)∞ − Ew) + τwbr∞. We multiply both sides of eq. (14) by V and integrate over the interval (−∞, Vs], as- suming that p∞(V ) tends sufficiently quickly toward zero for V → −∞ (Brunel 2000; Brunel et al. 2003), to obtain an equation which relates the steady-state spike rate and mean membrane voltage, (cid:104)(cid:104)V (cid:105)∞−EL +∆T (cid:68) exp (cid:16) V −VT (cid:17)(cid:69) ∆T (cid:105) µa − gL r∞ = ∞ /C , ∆V + τwb/C (18) where µa := µ − a((cid:104)V (cid:105)∞−Ew)/C, ∆V := Vs − Vr (here and in the following) and (cid:104)·(cid:105)∞ denotes the average with 3 A + a (cid:115)(cid:18) (cid:19)2  , respect to the density p∞(V ). The spike rate r∞ is given by eq. (18) only for nonnegative values of the numerator (i.e., µa − gL[. . . ]/C ≥ 0); otherwise, r∞ is defined to be zero. For simplicity, the refractory period Tref is omitted here. Note, that the steady-state spike rate for Tref (cid:54)= 0 can be calculated as r∞/(1 + r∞Tref ). We cannot express p∞(V ) explicitly and thus the expressions for the aver- ages with respect to p∞(V ) in eq. (18) are not known. However, in the case gL = 0, which simplifies the aEIF model to the aPIF model, an explicit expression for (cid:104)V (cid:105)∞ can be derived. We multiply eq. (14) by V 2 and integrate over [−∞, Vs] on both sides (assuming again that p∞(V ) quickly tends to zero for V → −∞) to obtain (cid:104)V (cid:105)∞ = 1 2a Vs + Vr 2 − A − a Vs + Vr 2 + B (19) where A = µC + aEw and B = 2aσ2C[1 + τwb/(C∆V )]. I-O curve The I-O curve is specified by the spike rate as a function of input strength. Here we consider two types of I-O curves: a time-varying (adapting) I-O curve and the steady-state I-O curve. In particular, we obtain the adapting I-O curve as the instantaneous spike rate response to a sustained input step (with a small baseline input) as a function of step size. This curve changes (adapts) over time and it eventually converges to the steady-state I-O curve. As arguments of these (adapting and steady-state) I-O functions we consider presynaptic spike rates (Figs. 2C, 4B and eq. (38)), input mean and standard deviation1 (Figs. 2D, 4B and eq. (36)) and input mean for fixed values of input standard deviation (Fig. 8A). ISI distribution We calculate the ISI distribution for an aEIF neuron which has reached a steady-state spike rate r∞ := limt→∞ r(t) by solving the so-called first passage time problem (Risken 1996; Tuckwell 1988). Consider an ini- tial condition where the neuron has just emitted a spike and the refractory period has passed. That is, the mem- brane voltage is at the reset value Vr and the adap- tation current, which we have replaced by its trial av- erage (see above), takes the value ¯w0, where ¯w0 will be determined self-consistently (see below). In each of N (simultaneous) trials, we follow the dynamics of the neuron given by dVi/dt = [Iion(Vi; ¯w) + Isyn(t)]/C, its mem- brane voltage crosses the value Vs and record that spike time Ti. The set of times Ti + Tref then gives the ISI distribution. Finally, we determine ¯w0 by imposing that the mean ISI matches with the known steady-state spike i=1 Ti + Tref = r−1∞ . According to this calculation scheme, the ISI distribution can be obtained in the limit N → ∞ by solving the Fokker-Planck system eqs. (8) -- (9) with mean adaptation current governed by d ¯w/dt = [a(1/N(cid:80)N rate, i.e., 1/N(cid:80)N i=1 Vi − Ew) − ¯w]/τw, until τw d ¯w dt = a((cid:104)V (cid:105)p(V,t) − Ew) − ¯w, (20) 1Note that because of two arguments we obtain a surface instead of a curve in this case. subject to the boundary conditions (12) and initial con- ditions p(V, 0) = δ(V − Vr), ¯w(0) = ¯w0. Note that the re-injection condition eq. (11) is omitted (see also the dif- ference between eqs. (10) and (20)) because here each trial i ends once Vi(t) crosses the value Vs. The ISI distribu- tion is given by the probability flux at Vs (Tuckwell 1988; Ostojic 2011), taking into account the refractory period, pISI(T ) = q(Vs, T − Tref ) 0 for T ≥ Tref for T < Tref . (21) (cid:40) Finally, ¯w0 is determined self-consistently by requiring (cid:104)T(cid:105)pISI = r−1∞ . The coefficient of variation (CV) of ISIs is then calculated as (cid:113)(cid:104)T 2(cid:105)pISI − (cid:104)T(cid:105)2 pISI CV := (cid:104)T(cid:105)pISI . (22) An ISI CV value of 0 indicates regular, clock-like spiking, whereas for spike times generated by a Poisson process the ISI CV assumes a value of 1. For a demonstration of the ISI calculation scheme described above see Fig. 1. The results based on the Fokker-Planck equation and numerical simulations of the aEIF model with fluctuating input are presented for an increased subthreshold and spike-triggered adaptation current in separation. ISI CV for the aPIF model To calculate the ISI CV we need the first two ISI moments, cf. eq. (22). The mean ISI for the aPIF neuron model is simply calculated by the inverse of the steady-state spike rate, cf. eq. (18), derived in the previous section, µ. Combining eqs. (22),(23) and (25) the ISI CV reads (cid:112)σ2∆V /µa − τ 2 CV = wb2/C 2 − 2τwb∆V /C . (26) ∆V + τwb/C Neuronal network To investigate the effects of recurrent (inhibitory) synap- tic inputs on the neuronal response properties (spike rates and ISIs), we consider a network instead of a single neuron, consisting of NE excitatory and NI inhibitory aEIF neu- rons (with separate parameter sets). The two populations are recurrently coupled in the following way (see Fig. 4A). Each excitatory neuron receives inputs from K extEE exter- nal excitatory neurons which produce instantaneous PSPs of magnitude J extEE with Poisson rate rextEE (t). Analogously, each inhibitory neuron receives inputs from K extIE exter- nal excitatory neurons producing instantaneous PSPs of magnitude J extIE with Poisson rate rextIE (t). In addition, each excitatory neuron receives inputs from K recEI randomly selected inhibitory neurons of the network with synap- tic strength (i.e., instantaneous PSP magnitude) J recEI and each inhibitory neuron receives inputs from K recIE randomly selected excitatory neurons of the network with synaptic strength J recIE . This network setup was chosen to exam- ine the effects caused by recurrent inhibition and compare them to the effects produced by external inhibition for single neurons described above. To reduce the parameter space, recurrent connections within the two populations in the network were therefore omitted. The total synaptic current for each neuron of the network can be described using eq. (4), where the parameters µ(t) and σ(t) for ex- citatory neurons are given by (cid:104)T(cid:105)pISI = r−1∞ = ∆V + τwb/C µa , (23) µ(t) = J extEE K extEE rextEE (t) + J recEI K recEI rpopI σ(t)2 = (J extEE )2K extEE rextEE (t) + (J recEI )2 K recEI rpopI (t), (t) (27) (28) where we consider µa > 0 (here and in the following). We approximate the second ISI moment by solving the first passage time problem for the Langevin equation exp(−t/τw) + ση(t), (24) dV dt = µa − ¯w0 C with initial membrane voltage Vr and boundary voltage Vs. That is, we replace (cid:104)V (cid:105)p(V,t) by its steady-state value (cid:104)V (cid:105)∞ in eq. (20), which is justified by large τw (as already assumed). The first passage time density (which is equiv- alent to pISI) and the associated first two moments for this type of Langevin equation can be calculated as power se- ries in the limit of small ¯w0 (Urdapilleta 2011). ¯w0 is then determined self-consistently by imposing eq. (23). Here we approximate the second ISI moment by using only the most dominant term of the power series, which yields (the zeroth order approximation) (Urdapilleta 2011), (cid:104)T 2(cid:105)pISI = σ2∆V + µa∆V 2 µ3 a . (25) Including terms of higher order leads to a complicated expression for (cid:104)T 2(cid:105)pISI which has to be evaluated numer- ically. We additionally considered the first order term (not shown) and compared the results of both approxima- tions (see Results). Effectively, the approximation above, eq. (25), is valid for small levels of spike-triggered adapta- tion current and mean input, since ¯w0 increases with b and 4 and for inhibitory neurons, µ(t) = J extIE K extIE rextIE (t) + J recIE K recIE rpopE σ(t)2 = (J extIE )2K extIE rextIE (t) + (J recIE )2 K recIE rpopE (t), (t) (29) (30) (t) and rpopI (Brunel 2000; Augustin et al. 2013). rpopE (t) are the spike rates of the excitatory and inhibitory neurons of the network, respectively. Here we consider large populations of neurons instead of a large number of trials. In fact, averaging over a large number of trials in this setting is equivalent to averaging over large populations due to the random and sparse connectivity. In the limit NE , NI → ∞ we obtain a system two coupled Fokker-Planck equations, one for the excitatory popula- tion, described by eqs. (8) -- (13),(27),(28), and one for the inhibitory population, given by eqs. (8) -- (13),(29),(30). Note that r(t) in eqs. (10) and (13) is replaced by the spike rates of the excitatory and inhibitory popula- tions, rpopE (t), respectively. We solve this system to obtain the steady-state spike rate for each population, rpopE,∞ and rpopI,∞. Once these quantities are known, we calculate the ISI distribution, cf. eq. (21), for the excitatory population (i.e. for any neuron of that population) as described above, using eqs. (27) -- (28) for the (steady-state) moments of the synaptic current. The neuron model parameter values were as above for the single neuron, with a = 0.015 mS/cm2, b = 0.1 µA/cm2 for excitatory neurons and a = b = 0 for inhibitory (t) and rpopI neurons, since adaptation was found to be weak in fast- spiking interneurons compared to pyramidal neurons (La Camera et al. 2006). The network parameter values were J extEE = J extIE = 0.15 mV, K extEE = K extIE = 800, constant rextEE ∈ [0, 80] Hz, J recEI ∈ [−0.75, −0.45] mV, K recEI = 100, constant rextIE ∈ [6, 14] Hz, J recIE ∈ [0.05, 0.2] mV and K recIE = 400. Numerical solution We treated the Fokker-Planck equations for the aPIF model analytically. In case of the aEIF model, we solved these equations forward in time using a first-order finite volume method on a non-uniform grid with 512 grid [−200 mV, Vs] and the implicit points in the interval Euler integration method with a time step of 0.1 ms for the temporal domain. For more details on the numerical solution, we refer to (Augustin et al. 2013). Detailed conductance-based neuron model For validation purposes we used a biophysical Hodgkin- Huxley-type neuron model with different types of slow K+- currents. The membrane voltage V of this neuron model obeys the current balance equation C dV dt = I − IL − INa − IK − ICa − IKs, (31) where C = 1 µF/cm2 is the membrane capacitance and I denotes the injected current. The ionic currents consist of a leak current, IL = gL(V − EL), a spike-generating Na+-current, INa = gNa(V )(V − ENa), a delayed recti- fier K+-current, IK = gK(V )(V − EK), a high-threshold Ca2+-current, ICa = gCa(V )(V − ECa), and a slow K+- current IKs. gx denote the conductances of the respec- tive ion channels and Ex are the reversal potentials. We separately considered three types of slow K+-current: a Ca2+-activated current (IKs ≡ IKCa) which is associated with the slow after-hyperpolarization following a burst of spikes (Brown and Griffith 1983), a Na+-activated cur- rent (IKs ≡ IKNa) (Schwindt et al. 1989), and the voltage- dependent muscarine-sensitive (M-type) current (IKs ≡ IM) (Brown and Adams 1980). The leak current depends linearly on the membrane potential. All other ionic cur- rents depend on V in a non-linear way as described by the Hodgkin-Huxley formalism. We adopted the somatic model from (Wang et al. 2003) and included the M-current with dynamics described (for the soma) by (Mainen and Sejnowski 1996). The conductances underlying the cur- rents INa, IK, ICa and IM are given by gNa = ¯gNam3∞h, gK = ¯gKn4, gCa = ¯gCas2∞ and gM = ¯gMu, respectively, with steady-state gating variables m∞ = αm/(αm + βm), αm = −0.4(V + 33)/(exp(−(V + 33)/10) − 1), βm = 16 exp(−(V + 58)/12) and s∞ = 1/[1 + exp(−(V + 20)/9)]. The dynamic gating variables x ∈ h, n, u are governed by dx dt = αx(1 − x) − βxx, (32) where αh = 0.28 exp(−(V + 50)/10), βh = 4/[1 + exp(−(V + 20)/10)], αn = −0.04(V + 34)/[exp(−(V + 34)/10) − 1], βn = 0.5 exp(−(V + 44)/25), αu = 3.209 · 10−4(V + 30)/[1 − exp(−(V + 30)/9)] and βu = −3.209 · 10−4(V + 30)/[1− exp((V + 30)/9)]. The channel opening and closing rates αx and βx are specified in ms−1 and the membrane voltage V in the equations above is replaced by d[Ca] its value in mV. The conductance for the Ca2+-activated slow K+-current IKCa is given by gKCa = ¯gKCa[Ca]/([Ca]+ κ), where the intracellular Ca2+-concentration [Ca] satis- fies = −αCaICa − [Ca] τCa (33) with αCa = 6.67 · 10−4 µM cm2/(µA ms), τCa = 240 ms and κ = 0.03 mM. The conductance for the Na+ -- activated slow K+-current IKNa is described by gKNa = ¯gKNa0.37/(1 + (/[Na])3.5) where  = 38.7 mM and the intracellular Na+-concentration [Na] is governed by dt (cid:18) [Na]3 [Na]3 + ϑ3 − γ (cid:19) (34) d[Na] dt = −αNa − 3ϕ with αNa = 0.3 µM cm2/(µA ms), ϕ = 0.6 µM/ms, ϑ = 15 mM and γ = 0.132. We varied the peak conduc- tances of the three slow K+-currents IKCa, IKNa, IM in the ranges ¯gKCa ∈ [2, 8] mS/cm2, ¯gKNa ∈ [2, 8] mS/cm2 (Wang et al. 2003) and ¯gM ∈ [0.1, 0.4] mS/cm2 (Mainen and Sejnowski 1996). The remaining parameter values were C = 1 µF/cm2, gL = 0.1 mS/cm2, EL = −65 mV, ENa = 55 mV, EK = −80 mV, ECa = 120 mV (Wang et al. 2003). The differences of the slow K+-currents (IKCa, IKNa and IM) is effectively expressed by their steady-state volt- age dependence and time constants. Therefore, we fur- ther considered a range of biologically plausible steady- state conductance-voltage relationships and timescales us- ing the generic description of a slow K+-current, IKs = ¯gKs ω(V )(V − EK), with peak conductance ¯gKs and gating variable ω(V ) given by = ω∞(V ) − ω, dω dt τω (35) where ω∞(V ) = 1/[1 + exp(−(V − α)/β)]. The shape of the steady-state curve ω∞(V ) was changed by the pa- rameters α ∈ [−40, −10] mV (half-activation voltage), β ∈ [6, 12] mV (inverse steepness) and the time con- stant τω was varied in [100, 300] ms. The model equations were solved using a second order Runge-Kutta integration method with a time step of 10 µs. To examine the effects of slow K+-currents on the I-O curve and ISI variability for noisy input, we additionally considered the synaptic current described by eq. (4) for i.e., we used I ≡ Isyn in the detailed neuron model, eq. (31). Subthreshold and spike-triggered components of biophysical slow K+-currents To assess how the relative levels of subthreshold adapta- tion conductance (parameter a) and spike-triggered adap- tation current increments (parameter b) in the aEIF model reflect different types of slow K+-currents, we quantified their subthreshold and spike-triggered components using the detailed conductance-based neuron model. First, we fit the steady-state adaptation current w∞ = a(V − Ew) from the aEIF model to the respective K+-current IKs of the Hodgkin-Huxley-type model in steady-state over a range of subthreshold values for the membrane voltage, V ∈ [−70, −60] mV. Thereby we obtained an estimate for a. In the second step, we measured the absolute and rel- ative change of IKs elicited by one spike. This was done 5 s s s ) − IKs(tpre by injecting a slowly increasing current ramp into the de- tailed model neuron and measuring IKs just before and after the first spike that occurred. Specifically, the ab- solute change of current caused by a spike was given by ∆IKs := IKs(tpost ), where the time points tpre and tpost were defined by the times at which the mem- brane potential crosses a value close to threshold (we chose −50 mV) during the upswing and downswing of the spike, respectively. ∆IKs provides an estimate for b. The relative change of K+-current was ∆I rel ). Here we only fitted the parameters a and b of the aEIF model. For an alternative fitting procedure which comprises all model parameters, we refer to (Brette and Gerstner 2005). Ks := ∆IKs/IKs(tpre s s Results Spike rate adaptation, gain and threshold modu- lation in single neurons We first examine the responses of single aEIF neurons with and without an adaptation current, receiving inputs from stochastically spiking presynaptic excitatory and in- hibitory neurons. The compound effect of the individual synaptic inputs is represented by an ongoing fluctuating input current whose mean and standard deviation depend on the synaptic strengths and spike rates of the presy- naptic cells (cf. eqs. (4) -- (6) in Materials and Methods and Fig. 2A). The neurons naturally respond to a sudden increase in spike rate of the presynaptic neurons (an in- put step) with an abrupt increase in spike rate and mean membrane voltage, see Fig. 2B. Without an adaptation current, both quantities remain unchanged after that in- crease. In case of a purely subthreshold adaptation cur- rent (a > 0, b = 0 in the aEIF model) which is present already in absence of spiking, the rapid increase of mean membrane voltage causes the mean adaptation current to build up slowly, which in turn leads to a gradual decrease in spike rate and mean membrane voltage. Note that the mean membrane voltage is decreased in the absence of spiking (before the increase of input) compared to the neu- ron without adaptation. In case of a purely spike-triggered adaptation current (a = 0, b > 0 in the aEIF model), the sudden increase in spike rate leads to an increase of mean adaptation current, which again causes the spike rate and mean membrane voltage to decrease gradually. The adapting I-O curve of neurons with and without an adaptation current, that is, the time-varying spike rate response to a step in presynaptic spike rates as a function of the step size, is shown in Fig. 2C. Interestingly, the two types of adaptation current affect the spike rate response in different ways. A subthreshold adaptation current shifts the I-O curve subtractively and thus increases the thresh- old for spiking. In addition, it decreases the response gain for low (output) spike rates. If the adaptation current is driven by spikes on the other hand, the I-O curve changes divisively, that is, the response gain is reduced over the whole range of spike rate values but the response thresh- old remains unchanged. It can be recognized that for a given type of adaptation current the adapting I-O curve evaluated shortly after the input steps and the steady- state I-O curve are changed qualitatively in the same way. Thus, for the following parameter exploration and analyt- ical derivation we focus on (changes of) the steady-state I-O relationship. We next explore the effects of an adaptation current on the steady-state spike rate for a wide range of input statistics, that is, different values of the mean µ and the standard deviation σ of the fluctuating total synaptic in- put, see Fig. 2D. If excitatory and inhibitory inputs are approximately balanced, the standard deviation σ of the compound input is large compared to its mean µ. The spike rate increases with an increase of either µ or σ, or both. A subthreshold adaptation current increases the threshold for spiking in terms of µ as well as σ. A spike- triggered adaptation current however does not change the threshold for spiking but reduces the gain of the spike rate as a function of µ or σ. Thus, the differential effects of both types of adaptation current are robust across dif- ferent input configurations. Note that the I-O curve as a function of mean input µ changes additively for increased levels of standard deviation σ while its slope (i.e., gain) decreases, particularly for small values of µ. This can be recognized by the contour lines in Fig. 2D and is most prominent for increased subthreshold adaptation. Con- sequently, this type of adaptation current increases the sensitivity of the steady-state spike rate to noise intensity for low spike rates. In order to analytically demonstrate the differential ef- fects of subthreshold and spike-triggered adaptation cur- rents on the (steady-state) I-O curve, we consider the aPIF neuron model, which is obtained by neglecting the leak conductance (gL = 0) in the aEIF model. This allows to derive an explicit expression for the steady-state spike rate, µ − a((cid:104)V (cid:105)∞−Ew)/C r∞ = , ∆V + τwb/C (36) where the mean membrane voltage (cid:104)V (cid:105)∞ with respect to the steady-state distribution p∞(V ) is given by eq. (19) and ∆V := Vs − Vr is the difference between spike and reset voltage; r∞ = 0 for µ < a((cid:104)V (cid:105)∞ − Ew)/C (see Materials and Methods). Equation (36) mathematically demonstrates the subtractive component of the effect a subthreshold adaptation current (a > 0) produces when the mean membrane voltage is larger than the reversal potential Ew of the (K+) adaptation current. Taking the derivative of eq. (36) with respect to µ further reveals that an increase of a reduces the gain when the input fluctuations (σ) are large compared to the mean (µ). A spike-triggered adaptation current (b > 0) however produces a purely divisive effect which can be pronounced even for small current increments b if the adaptation timescale τw is large. Differential effects of adaptation currents on spik- ing variability We next investigate how adaptation currents affect ISIs for different input statistics. For that reason we calculate the distribution of times at which the membrane voltage of an aEIF neuron crosses the threshold Vs for the first time, which is equivalent to the distribution of ISIs (see Mate- rials and Methods). These ISI distributions are shown in Fig. 3A for neurons with different levels of subthreshold or spike-triggered adaptation and a given input. An increase of either type of adaptation current (via parameters a and b) naturally increases the mean ISI. Interestingly, while subthreshold adaptation leads to ISI distributions with 6 long tails, spike-triggered adaptation causes ISI distribu- tions with bulky shapes. These differential effects on the shape of the ISI distribution lead to opposite changes of the coefficient of variation (CV, cf. eq. (22)) which quan- tifies the variability of ISIs. An increase of subthreshold adaptation curent produces an increase of CV, whereas an increase of spike-triggered adaptation current leads to a decreased ISI variability. How these effects on the CV of ISIs depend on the statistics (µ and σ) of the fluctuating input is shown in Fig. 3B,C. With or without an adap- tation current, if the mean µ is large, that is, far above threshold, and the standard deviation σ is comparatively small, the neuronal dynamics is close to deterministic and the firing is almost periodic, hence the CV is small. In con- trast, if µ is close to the threshold and σ is large (enough), the ISI distribution will be broad as indicated by the large CV. A subthreshold adaptation current either leads to an increased CV or leaves the ISI variability unchanged. In case of a spike-triggered adaptation current the effect on the CV depends on the input statistics. This type of adap- tation current causes a decrease of the high ISI variability in the region (of the µ, σ-plane) where the mean input µ is small, and an increase of the low ISI variability for larger values of µ. We analytically derived an approximation of the ISI CV for the aPIF model, which emphasizes the opposite effects of the two types of adaptation current. It is obtained as CV = ∆V + τwb/C (37) (same as eq. (26)), where µa := µ − a[(cid:104)V (cid:105)∞−Ew]/C is the effective mean input which is again assumed to be positive and takes into account the counteracting subthreshold adaptation current. The steady-state mean membrane voltage (cid:104)V (cid:105)∞ is given by eq. (19) (see Materials and Methods). Equation (37) mathematically demonstrates that an increase of subthreshold adaptation curent (a > 0) causes an increase of CV as long as (cid:104)V (cid:105)∞ is larger than Ew, that is, the mean membrane voltage is not too hyperpolarized. An increase of spike-triggered adaptation current (b > 0) on the other hand leads to a reduction of ISI variability. Note that this approximation is only valid for small values of mean input (µ) and adap- tation current increment (b). It does not account for the increase of CV caused by spike-triggered adaptation for large levels µ, cf. Fig. 3C. Both (input dependent) effects of spike-triggered adaptation on the ISI variability can be captured by a refined approximation of the CV compared to eq. (37) (not shown, see Materials and Methods for an outline), which requires numerical evaluation. (cid:112)σ2∆V /µa − τ 2 wb2/C 2 − 2τwb∆V /C model parameters within each population and sparse ran- dom connectivity (see Materials and Methods). Fig. 4B shows how the steady-state I-O curve of excitatory neu- rons, i.e., the spike rate rpopE,∞ as a function of the external (input) spike rate rextEE , is changed by external excitation to the inhibitory neurons (via rextIE ) and by the strengths of the recurrent excitatory and inhibitory synapses (J recIE and J recEI ), respectively. An increase of external excitation to the inhibitory population (via rextIE ) changes the I-O curve subtractively, thus increasing the response thresh- old, while an increase of recurrent excitation to the in- hibitory neurons (via J recIE ) has a purely divisive effect, that is, the gain is reduced. On the other hand, an in- crease of recurrent inhibition to the excitatory neurons (via J recEI ) affects the I-O curve in both ways. We demonstrate these effects analytically for a network of perfect integrate-and-fire (PIF) model neurons (instead of aEIF neurons). That is, we disregard the adaptation current here for simplicity (a = b = 0), since it does not change the results qualitatively. An explicit expression for the steady-state spike rate of the excitatory neurons, rpopE,∞, can be derived using eq. (36) for all the neurons in the network with mean input µ given by eq. (27) for excitatory neurons and by eq. (29) for inhibitory neurons. We solve for rpopE,∞ self-consistently to obtain, rpopE,∞ = J extEE K extEE rextEE ∆V + J recEI K recEI J extIE K extIE rextIE ∆V 2 − J recIE K recIE J recEI K recEI . (38) The equation above states that rpopE,∞ is directly pro- portional to the strength of external excitation to the excitatory population, negatively proportional to the strength of external excitation to the inhibitory pop- ulation (since J recEI < 0) and inversely proportional to the strength of recurrent excitation, where all propor- tionalities include an offset. Eq. (38) clearly shows that the effect of external excitation to the inhibitory population is purely subtractive (since J recEI < 0), the effect of recurrent excitation (to the inhibitory popu- lation) is purely divisive, and the effect of recurrent inhibition (to the excitatory population) includes both components. For comparison, consider a single (non- adapting) PIF neuron receiving (external) excitatory and inhibitory input. Using eq. (36) with mean input µ given by eq. (5), the steady-state spike rate of this neuron reads r∞ = (JE KE rE + JIKIrI)/∆V . Thus, an increase of external inhibition affects the I-O curve of an excitatory neuron in the same way (subtractively) as an increase of external excitation to the inhibitory population within a recurrent network as described above. Differential effects of synaptic inhibition on I-O curves Here we examine how synaptic input received from a pop- ulation of inhibitory neurons affect gain and threshold of spiking. We consider that the neuron we monitor belongs to a population of excitatory neurons which are recurrently coupled to neurons from an inhibitory population, as de- picted in Fig. 4A: Each neuron of the network receives excitatory synaptic input from external neurons and ad- ditional synaptic input from a number of neurons of the other population. The specific choice of the monitored excitatory neuron does not matter because of identical Effects of synaptic inhibition on spiking variability We next investigate how inhibitory synaptic input changes the ISI variability of the neurons (from the excitatory pop- ulation) in the network described above. An increase of external excitation to the inhibitory neurons (via rextIE ), and the strengths of the recurrent synapses (J recIE and J recEI ) individually, leads to an increase of the mean ISI and an increased tail of the ISI distribution, as shown in Fig. 5A. Furthermore, an increase of rextIE or the magnitude of J recIE or J recEI , each causes the coefficient of variation of ISIs (CVpopE ) to increase, see Fig. 5B. Thus, an increase of in- hibition always leads to an increase of spiking variability. 7 . An increase of external excitation to the excitatory neu- rons (via rextEE ), on the other hand, leads to a decrease of CVpopE To demonstrate these effects analytically we derived CVpopE for a network of PIF model neurons using eqs. (26) -- (28), where we obtained the steady-state spike rate of the inhibitory neurons, rpopI,∞, analogously to rpopE,∞ (as de- scribed above). Below, we express CVpopE as a function of either rextIE , J recIE or J recEI , and lump together all other fixed parameters in a number of constants, (c1rextIE + c2)/(c3 − c4rextIE ) c5J recIE + c6 (c7(J recEI )2 − c8J recEI )/(c9 + c10J recEI ). CVpopE = (39) The constants c1, . . . , c10 in eq. (39) are non-negative func- tions of the fixed parameters. Clearly, an increase of rextIE or the magnitudes of J recIE and J recEI each produce an in- crease of CVpopE (since J recEI < 0). Considering a single PIF neuron receiving (external) excitatory and inhibitory input for comparison, we use eq. (37) with mean µ and standard deviation σ of the input given by eqs. (5) and (6), respectively, to express the CV as (cid:115) CV = J 2E KE rE + J 2IKIrI ∆V (JE KE rE + JIKIrI) . (40) The effect of Note that eq. (40) is only valid for positive mean input (JE KE rE + JIKIrI > 0). Again, ISI variability increases with inhibition. inhibition on spiking variability can be understood intuitively as follows. Inhibitory synaptic input reduces the mean total synaptic input µ and increases its standard deviation σ for the target neuron (population), which in turn causes an increase of ISI variability. Subthreshold and spike-triggered components of slow K+-currents Here we examine how the two types of an adaptation cur- rent in the aEIF model reflect different slow K+-currents in a detailed conductance-based neuron model. First, we consider three prominent slow K+-currents: a Ca2+- activated after-hyperpolarization current (IKCa), a Na+- activated current (IKNa) and the voltage-dependent M- current (IM). Fig. 6A shows how the conductances asso- ciated with these K+-currents depend on the membrane voltage in the steady state, compared to the steady-state spike-generating Na+-conductance. The threshold mem- brane voltage at which a spike is elicited in response to a slowly increasing input current is primarily determined by the conductance-voltage relationship for Na+. The thresh- old value lies in the interval where this curve has a positive slope (the precise value depends on the peak conductances of all currents and on the input). The curve gNa,∞(V ) thus indicates the subthreshold and suprathreshold mem- brane voltage ranges. In the subthreshold voltage range the conductance gKCa,∞ is almost zero, while the conduc- tances gKNa,∞ and gM,∞ reach significant values close to the voltage threshold. Thus, the curves in Fig. 6A indicate that IKCa is activated by spikes, while IM and particularly IKNa can be increased in the absence of spiking. The results of the fitting procedure in Fig. 6B,C show the absolute and relative amounts of current triggered by a 8 spike versus its subthreshold level quantified by the voltage independent conductance a. IKCa has a dominant spike- triggered component as expected, while IKNa shows a very small increment caused by a spike compared to the sub- threshold component. IM, on the other hand, shows signif- icant levels of both components. Note, however, that the amount of IM elicited by a spike is smaller compared to the level of IM that can be caused by subthreshold mem- brane depolarization without spiking (since ∆I rel Ks < 1 for IKs ≡ IM, see Fig. 6C). relationship for K+, We further considered a range of biologically plausible slow K+-currents. That is, we varied the steady-state gKs,∞(V ), conductance-voltage within a realistic range, as shown in Fig. 7A, and quantified the subthreshold and spike-triggered com- ponents for each of these K+-currents, see Fig. 7B,C. The value of subthreshold conductance a naturally increases with the fraction of K+-conductance present at subthreshold voltage values. For the quantification of spike-triggered current increments we also considered different K+ time constants τω. The absolute value of current increment ∆IKs decreases with increasing τω and changes only slightly with changes of the shape of the conductance-voltage curve gKs,∞(V ) (via the parameters α, β). However, the current increment caused by a spike relative to the amount of current already present in the absence of spiking (∆I rel Ks) is strongly determined by gKs(V ). ∆I rel increases with an in- Ks crease of half-activation voltage (parameter α), steepness (via parameter β) and with decreasing time constant (τω). Effects of slow K+-currents on I-O curve and ISI variability Here we examine how the different types of slow K+- current affect the I-O curve and spiking variability of uncoupled conductance-based model neurons subject to noisy inputs and compare the effects to those caused by subthreshold and spike-triggered adaptation in aEIF neu- rons. Without a slow K+-current, the spike rate I-O curve does not change over time, see Fig. 8A. An increase of IKCa has a purely divisive effect on the I-O curve while an in- crease of IM changes this curve in a mostly subtractive and slightly divisive way. For both types of slow K+-current the adapting spike rates reach their steady-state values in less than 500 ms. These effects are consistent with our results based on the aEIF model, given that IKCa pre- dominantly depends on spikes and IM includes both, sub- threhold as well as spike-triggered, components (Fig. 6B). In case of increased IKNa, on the other hand, the steady- state I-O curve is significantly altered in both ways (sub- tractively and divisively), and the spike rates adapts very slowly, that is, steady-state rates are reached after several seconds. At first sight, this seems contradictory to the ef- fect predicted above for subthreshold adaptation, consid- ering that the amount of IKNa triggered by a spike is small compared to its subthreshold level. Since the timescale of IKNa is very large (Fig. 8A and (Wang et al. 2003)) even a small spike-triggered component leads to a significant divi- sive change of the steady-state I-O curve, cf. eq. (36). This divisive effect is caused by K+-current building up slowly because of small current increments triggered repeatedly by repetitive spiking and very slow decay between spikes due to the large timescale of the current. Considering ISI variability, an increase of IKCa reduces the CV for small values of mean input µ and increases the CV for larger values of µ, see Fig. 8B. An increase of each of the other slow K+-currents, IKNa and IM, leads to an increase of ISI CV in general. These effects are consistent with those caused by subthreshold and spike- triggered adaptation currents in the aEIF model, consid- ering the subthreshold and spike-triggered components of IKCa, IKNa and IM, respectively (Fig. 6). Thus, the re- sults from the detailed conductance-based neuron model are in agreement with the results based on the adaptive integrate-and-fire models presented above. Discussion In this study, we have systematically examined how adap- tation currents and synaptic inhibition modulate the threshold and gain of spiking as well as ISI variability in response to fluctuating inputs resulting from stochastic synaptic events. Based on a simple neuron model with subthreshold and spike-triggered adaptation components we used analytical and numerical tools to describe spike rates and ISIs for a wide range of input statistics. We then measured subthreshold and spike-triggered compo- nents of different types of slow K+-currents using detailed conductance-based model neurons and we validated our (analytical) results from the simple neuron model by nu- merical simulations of the detailed model. We have shown that a purely subthreshold voltage- dependent adaptation current increases the threshold for spiking and reduces the gain at low spike rates in the pres- ence of input fluctuations. This type of current produces a long-tailed ISI distribution and thus leads to an increase of variability for a broad range of input statistics. A spike- triggered adaptation current, on the other hand, causes a divisive change of the I-O curve, thereby reducing the re- sponse gain but leaving the response threshold unaffected, irrespective of the input noise intensity. This type of cur- rent decreases the ISI CV for fluctuation-dominated inputs but increases the CV when the mean input is strong, i.e., it reduces the sensitivity of spiking variability to the mean input. For comparison, an increase of external inhibition leads to a subtractive shift of the I-O curve while an in- crease of recurrent inhibition changes it divisively. The ISI variability, however, is increased by both types of synaptic inhibition. We have further demonstrated that the Ca2+-activated after-hyperpolarization K+-current is effectively captured by a simple description based on spike-triggered incre- ments, while the muscarine-sensitive and Na+-activated K+-currents, respectively, have dominant subthreshold components. Despite its small spike-triggered component, the Na+-dependent K+-current also substantially affects the neuronal gain, due to its large timescale. Methodological aspects Our approach involves the diffusion approximation and Fokker-Planck equation, both of which have been widely applied to analyze the spike rates of scalar IF type neurons in a noisy setting, see e.g. (Amit and Brunel 1997; Brunel 2000; Fourcaud-Trocm´e et al. 2003; Burkitt 2006; Roxin et al. 2011). Our assumption of separated timescales be- tween slow adaptation and fast membrane voltage dynam- ics has also been frequently used in such a setting (Brunel et al. 2003; La Camera et al. 2004; Gigante et al. 2007b; Richardson 2009; Augustin et al. 2013). While most of these previous studies concentrated on spike rate dynam- ics, here we focused on asynchronous (non-oscillatory) activity. To examine ISI distributions we extended the method described previously for scalar IF models, which is based on the first passage time problem (Tuckwell 1988; Ostojic 2011), to the aEIF model, accounting for the dy- namics of the adaptation current between spikes. Further- more, we analytically derived an expression for the steady- state spike rate (i.e., steady-state I-O relationship) based on (Brunel et al. 2003) and an approximation of the ISI CV using recent results from (Urdapilleta 2011) for the per- fect IF model with two types of adaptation currents (aPIF model). The I-O functions we calculated can be used to re- late (adaptive) spiking neuron models to linear-nonlinear cascade models, which describe the instantaneous spike rate of a neuron by applying to the stimulus signal suc- cessively a linear temporal filter and a static nonlinear function (Ostojic and Brunel 2011). Such cascade models have proven valuable for studying how sensory inputs are mapped to neuronal activity (see, e.g., (Schwartz et al. 2006; Pillow et al. 2008)). It is worth noting that our approach further allows to easily calculate the power spectrum P and (normalized) autocorrelation function A of the neuronal spike train once the ISI distribution has been obtained, via the relation P(ω) = A(ω) = r∞Re , (41) (cid:18) 1 + pISI(ω) 1 − pISI(ω) (cid:19) to memoryless (so-called renewal) where A and pISI denote the Fourier transforms of the autocorrelation function and ISI distribution, respec- tively, see (Gerstner and Kistler 2002). Eq. (41) strictly applies stochastic processes and an adaptation mechanism usually leads to a violation of this requirement for a model neuron subject to fluctuating input. Here we have derived a renewal process (Vi(t), ¯w(t)) from the original non-renewal process (Vi(t), wi(t)) by averaging the adaptation current and self-consistently determining its reset value (see section ISI distribution in Materials and Methods). An alternative approach that allows for the application of the above relationship eq. (41) to adapting model neurons has recently been described in (Naud and Gerstner 2012). Modulation of spike rate threshold and gain Purely subtractive and divisive changes of the I-O curve by subthreshold and spike-triggered adaptation, respectively, have previously been shown for model neurons considering constant current inputs but neglecting input fluctuations (Prescott and Sejnowski 2008; Ladenbauer et al. 2012). These theoretical results describe the effects shown in re- cent in-vitro experiments which involved blocking the low- threshold M current and a Ca2+-activated K+-current sep- arately (Deemyad et al. 2012) (Fig. 3); see also (Alaburda et al. 2002) (Fig. 3), (Smith et al. 2002) and (Miles et al. 2005) (Fig. 1) for experimental evidence of either effect. Here we have shown that a subthreshold adaptation cur- rent also causes a reduction of response gain (in addition to an increase of response threshold) when the fluctuations of the input are strong compared to its mean. On the other 9 hand, a spike-triggered adaptation current decreases the response gain over the whole input range, irrespective of the level of input fluctuations. These results apply for adapting as well as the adapted (steady) states2. When considering the onset I-O curve, i.e., the immediate re- sponse to a sudden increase of input, an increased level of spike-triggered adaptation current due to pre-adaptation has been shown to produce a rather subtractive change (Benda et al. 2010). This, however, does not contradict our results. On the contrary, either type of adaptation current (subthreshold or spike-triggered) naturally leads to a subtractive change of the onset I-O curve for neurons which are pre-adapted to an increased input (not shown). Notably, when considering conductance based noisy synaptic inputs, an increase in balanced synaptic back- ground activity can also reduce the spike rate gain (Chance et al. 2002; Burkitt et al. 2003) and external inhibition can reduce the gain and increase the response treshold at the same time (Mitchell and Silver 2003). This means, the response gain can change due to external inputs that are independent of the activity of the target neuron, which can be understood as follows. An increase of noisy (ex- citatory or inhibitory) synaptic conductance leads to an increase of total membrane conductance, which causes a purely subtractive change of the I-O curve, and an increase in synaptic current noise, which causes an additive change of the I-O curve and decreases its slope (particularly for small input strengths) (Chance et al. 2002) (Fig. 3). Both effects combined lead to the observed change of response gain. The two separate components are included in our re- sults. An increase of membrane conductance (represented by gL in the aEIF model) subtracts from the spike rate response, see eq. (18), and the abovementioned effects of an increase of noise intensity σ have been described in the section Results (see Fig. 2D). Modulation of response gain is an important phe- nomenon, particularly in sensory neurons, because neu- ronal sensitivity to changes in the input is amplified or downscaled without changing input selectivity. A spike- dependent adaptation current thus represents a cellular mechanism by which this is achieved. For example, neu- ronal response gain increases during selective attention (McAdams and Maunsell 1999). It has been shown in- vivo that the neuromodulator acetylcholine (ACh) con- tributes substantially to attentional upregulation of spike rates (Herrero et al. 2008). Cholinergic changes of neu- ronal excitability and response gain (Soma et al. 2012) in turn are likely produced via downregulation of slow K+- currents (Madison et al. 1987; McCormick 1992; Sripati and Johnson 2006). Together with our results, these ob- servations suggest that excitability and response gain of cortical neurons are controlled by neuromodulatory sub- stances through (de)activation of subthreshold and spike- triggered K+-currents, respectively. We have further shown that external inhibitory synaptic inputs change the I-O curve subtractively, which is consis- tent with the results of a previous numerical study using a conductance based neuron model without consideration of noise (Capaday 2002). Recurrent synaptic (feedback) 2Note that in case of a very large adaptation timescale a (small) spike-triggered adaptation current has a negligible effect on the adapting I-O curve, evaluated shortly after the input steps, but a significant effect on the steady-state I-O curve (see Fig. 8A). inhibition, which is a function of the neuron's spike rate, on the other hand, reduces the response gain. This is in agreement with the results obtained by (Sutherland et al. 2009) based on IF type neurons subject to noisy inputs. Recent in-vivo recordings from mouse visual cortex have shown that distinct types of inhibitory neurons produce these differential effects (i.e., subtractive and divisive changes of I-O curves) at their target neurons (Wilson et al. 2012). Functional connectivity analysis suggests that the inhibitory neurons which changed the I-O curve of their target neurons subtractively were less likely connected recurrently to the recorded targets than the inhibitory neurons which changed their targets' responses divisively (Wilson et al. 2012). Applying our results based on the simple network model the observed differential effects caused by the two types of inhibitory cells can thus be explained by their patterns of connectivity with the target cells. Effects on ISI variability We have shown that a spike-triggered adaptation current reduces high ISI variability at low spike rates (when input fluctuations are strong compared to the mean) and in- creases low ISI variability at high spike rates (caused by a large mean input). This result is in agreement with a pre- vious numerical simulation study (Liu and Wang 2001) but seems to disagree with other theoretical work (Wang 1998; Prescott and Sejnowski 2008; Schwalger et al. 2010) at first sight. Wang (1998) and later Prescott & Sejnowski (2008) showed that spike-driven adaptation reduces the ISI CV at low spike rates but they did not find an increase of ISI CV at higher spike rates in their simulation studies. The reason for this is that the ISI CVs of adapting and non- adapting neurons were compared at equal spike rates (i.e., at equal mean ISIs) but different input statistics. That is, the input to the adapting neurons was adjusted to com- pensate for the change of spike rate (or mean ISI) caused by the adaptation currents. Increasing the mean input to the adapting neurons to achieve equal mean ISIs, however, decreases its ISI CV (cf. eq. (37)). Here we compare the ISI statistics across different neurons for equal inputs. On the other hand, Schwalger et al. (2010) analyzed the ISI statistics of perfect IF model neurons with spike-triggered adaptation and found that this type of adaptation always leads to an increase of ISI CV in response to a noisy in- put current. Their approach is similar to the one presented here but differs in that the dynamics of the adaptation cur- rent was neglected in (Schwalger et al. 2010), see Fig. 1B (bottom panel) for a visualization of that difference. As- suming a stationary adaptation current leads to a reduced effective mean input to the neuron, leaving the input vari- ance unchanged, which always causes increased ISI vari- ability, cf. eq. (37). Together with theoretical work show- ing that a spike-dependent adaptation current causes neg- ative serial ISI correlations (Prescott and Sejnowski 2008; Farkhooi et al. 2011) our results suggest that spike rate coding is improved by such a current for low-frequency in- puts (Prescott and Sejnowski 2008; Farkhooi et al. 2011). In contrast, an adaptation current which is predomi- nantly driven by the subthreshold membrane voltage usu- ally leads to an increase of ISI CV, as we have demon- strated. This seems to be not consistent with a previ- ous study (Prescott and Sejnowski 2008) where subthresh- 10 old adaptation was found to produce a small decrease of ISI variability. The apparent discrepancy is caused by differences in the presentation of the data: Prescott & Sejnowski (2008) compared the ISI CVs for equal spike rates as explained above. That is, the mean input was adjusted to obtain equal mean ISIs for adapting and non-adapting neurons but the input variance remained unchanged. However, increasing the mean input (µ in eq. (37)) to the adapting neuron counteracts the effect of subthreshold adaptation on the effective mean input (µa in eq. (37)). Consequently, one cannot observe an increased ISI CV in neurons with subthreshold adaptation currents when the mean input to these neurons is increased. Note that our results do not contradict those in (Prescott and Sejnowski 2008), but instead reveal that an increase of a subthreshold adaptation current always causes an increase of ISI CV for given input statistics and an increase of a spike-dependent adaptation current leads to an increase of ISI CV if the mean input is large. Finally, we have shown that an increase in synaptic in- hibition increases the ISI variability, regardless of whether this inhibition originates from an external population of neurons or from recurrently coupled ones. An intuitive ex- planation for this effect is that increased inhibitory input reduces the mean input but increases the input variance, see eqs. (5) -- (6). The reason why recurrent synaptic inhi- bition and spike-triggered adaptation change the ISI vari- ability in opposite ways in a fluctuation-dominated input regime could be the different timescales. Synaptic inhibi- tion usually acts on a much faster timescale than adapta- tion currents whose time constants range from about one hundred milliseconds to seconds. Thus, recurrent synaptic inhibition in contrast to spike-triggered adaptation cannot provide a memory trace of past spiking activity (over a duration of several ISIs) that could shape the ISI distri- bution. Notably, our results on ISIs in a network setting strictly apply to networks in asynchronous states. Recur- rent synaptic inhibition, however, can also mediate oscil- latory activity (Brunel 2000; Brunel et al. 2003; Isaacson and Scanziani 2011; Augustin et al. 2013) where the vari- ability of ISIs might be affected differently. Acknowledgements Augustin M, Ladenbauer J, Obermayer K. How adaptation shapes spike rate oscillations in recurrent neuronal networks. Front Comput Neurosci 7: 1 -- 11, 2013. Badel L, Lefort S, Brette R, Petersen CCH, Ger- stner W, Richardson MJE. Dynamic I-V curves are reliable predictors of naturalistic pyramidal-neuron voltage traces. J Neurophysiol 99: 656 -- 666, 2008. Benda J, Herz AVM. A universal model for spike- frequency adaptation. Neural Comput 15: 2523 -- 2564, 2003. Benda J, Longtin A, Maler L. Spike-frequency adap- tation separates transient communication signals from background oscillations. J Neurosci 25: 2312 -- 2321, 2005. Benda J, Maler L, Longtin A. Linear versus nonlinear signal transmission in neuron models with adaptation currents or dynamic thresholds. J Neurophysiol 104: 2806 -- 2820, 2010. Brette R, Gerstner W. Adaptive exponential integrate- and-fire model as an effective description of neuronal activity. J Neurophysiol 94: 3637 -- 3642, 2005. Brown DA, Adams PR. Muscarinic suppression of a novel voltage-sensitive K+ current in a vertebrate neu- rone. Nature 283: 673 -- 676, 1980. Brown DA, Griffith WH. Calcium-activated outward current in voltage-clamped hippocampal neurones of the guinea-pig. J Physiol 337: 287 -- 301, 1983. Brunel N. Dynamics of sparsely connected networks of excitatory and inhibitory spiking neurons. J Comput Neurosci 8: 183 -- 208, 2000. Brunel N, Hakim V, Richardson M. Firing-rate res- onance in a generalized integrate-and-fire neuron with subthreshold resonance. Phys Rev E 67: 051916, 2003. Burkitt AN. A review of the integrate-and-fire neuron model: I. Homogeneous synaptic input. Biol Cybern 95: 1 -- 19, 2006. This work was supported by DFG in the framework of collaborative research center SFB910. We thank Maziar Hashemi-Nezhad for helpful comments on the manuscript. Burkitt AN, Meffin H, Grayden DB. Study of neuronal gain in a conductance-based leaky integrate- and-fire neuron model with balanced excitatory and in- hibitory synaptic input. Biol Cybern 89: 119 -- 125, 2003. References Adams PR, Brown DA, Constanti A. Pharmacologi- cal inhibition of the M-current. J Physiol 332: 223 -- 262, 1982. Alaburda A, Perrier JF, Hounsgaard J. An M-like outward current regulates the excitability of spinal mo- toneurones in the adult turtle. J Physiol 540: 875 -- 881, 2002. Amit DJ, Brunel N. Model of global spontaneous ac- tivity and local structured activity during delay periods in the cerebral cortex. Cereb Cortex 7: 237 -- 252, 1997. Capaday C. A re-examination of the possibility of con- trolling the firing rate gain of neurons by balancing ex- citatory and inhibitory conductances. Exp Brain Res 143: 67 -- 77, 2002. Chance FS, Abbott LF, Reyes AD. Gain modulation from background synaptic input. Neuron 35: 773 -- 782, 2002. Compte A, Constantinidis C, Tegner J, Raghavachari S, Chafee MV, Goldman-Rakic PS, Wang XJ. Temporally irregular mnemonic persistent activity in prefrontal neurons of monkeys during a delayed response task. J Neurophysiol 90: 3441 -- 3454, 2003. 11 Deemyad T, Kroeger J, Chacron MJ. Sub- and suprathreshold adaptation currents have opposite ef- fects on frequency tuning. J Physiol 590: 4839 -- 4858, 2012. La Camera G, Rauch A, Luscher HR, Senn W, Fusi S. Minimal models of adapted neuronal response to in vivo-like input currents. Neural Comput 16: 2101 -- 2124, 2004. Destexhe A. Self-sustained asynchronous irregular states and up-down states in thalamic, cortical and thalam- ocortical networks of nonlinear integrate-and-fire neu- rons. J Comput Neurosci 27: 493 -- 506, 2009. La Camera G, Rauch A, Thurbon D, Luscher HR, Senn W, Fusi S. Multiple time scales of temporal response in pyramidal and fast spiking cortical neurons. J Neurophysiol 96: 3448 -- 3464, 2006. Destexhe A, Rudolph M, Par´e D. The high- conductance state of neocortical neurons in vivo. Nat Rev Neurosci 4: 739 -- 751, 2003. Destexhe A, Rudolph-Lilith M. Neuronal Noise. New York, USA: Springer, 2012. Ellis LD, Mehaffey WH, Harvey-Girard E, Turner RW, Maler L, Dunn RJ. SK channels provide a novel mechanism for the control of frequency tuning in elec- trosensory neurons. J Neurosci 27: 9491 -- 9502, 2007. Farkhooi F, Muller E, Nawrot M. Adaptation reduces variability of the neuronal population code. Phys Rev E 83: 050905, 2011. Fourcaud-Trocm´e N, Hansel D, van Vreeswijk C, Brunel N. How spike generation mechanisms deter- mine the neuronal response to fluctuating inputs. J Neurosci 23: 11628 -- 11640, 2003. Fuhrmann G, Markram H, Tsodyks M. Spike fre- quency adaptation and neocortical rhythms. J Neuro- physiol 88: 761 -- 770, 2002. Ladenbauer J, Augustin M, Shiau L, Obermayer K. Impact of adaptation currents on synchronization of coupled exponential integrate-and-fire neurons. PLoS Comput Biol 8: e1002478, 2012. Ladenbauer J, Lehnert J, Rankoohi H, Dahms T, Scholl E, Obermayer K. Adaptation controls syn- chrony and cluster states of coupled threshold-model neurons. Phys Rev E 88: 042713, 2013. Liu YH, Wang XJ. Spike-frequency adaptation of a gen- eralized leaky integrate-and-fire model neuron. J Com- put Neurosci 10: 25 -- 45, 2001. Madison DV, Lancaster B, Nicoll RA. Voltage clamp analysis of cholinergic action in the hippocampus. J Neurosci 7: 733 -- 741, 1987. Madison DV, Nicoll RA. Control of the repetitive dis- charge of rat CA1 pyramidal neurones in vitro. J Physiol 354: 319 -- 331, 1984. Maimon G, Assad JA. Beyond poisson: increased spike-time regularity across primate parietal cortex. Neuron 62: 426 -- 440, 2009. Gerstein GL, Mandelbrot B. Random walk models for the spike activity of a single neuron. Biophys J 4: 41 -- 68, 1964. Mainen ZF, Sejnowski TJ. Influence of dendritic struc- ture on firing pattern in model neocortical neurons. Na- ture 382: 363 -- 366, 1996. Gerstner W, Kistler WM. Spiking Neuron Models. Cambridge, UK: Cambridge University Press, 2002. Gigante G, Del Giudice P, Mattia M. Frequency- dependent response properties of adapting spiking neu- rons. Math Biosci 207: 336 -- 351, 2007a. Gigante G, Mattia M, Del Giudice P. Diverse population-bursting modes of adapting spiking neurons. Phys Rev Lett 98: 148101, 2007b. Herrero JL, Roberts MJ, Delicato LS, Gieselmann MA, Dayan P, Thiele A. Acetylcholine contributes through muscarinic receptors to attentional modulation in V1. Nature 454: 1110 -- 1114, 2008. Isaacson JS, Scanziani M. How inhibition shapes cor- tical activity. Neuron 72: 231 -- 243, 2011. Jolivet R, Schurmann F, Berger TK, Naud R, Ger- stner W, Roth A. The quantitative single-neuron modeling competition. Biol Cybern 99: 417 -- 426, 2008. Kim U, McCormick DA. Functional and ionic proper- ties of a slow afterhyperpolarization in ferret perigenic- ulate neurons in vitro. J Neurophysiol 80: 1222 -- 1235, 1998. McAdams CJ, Maunsell JHR. Effects of attention on orientation-tuning functions of single neurons in macaque cortical area V4. J Neurosci 19: 431 -- 441, 1999. McCormick DA. Neurotransmitter actions in the tha- lamus and cerebral cortex and their role in neuromod- ulation of thalamocortical activity. Progr Neurobiol 39: 337 -- 388, 1992. Miles GB, Dai Y, Brownstone RM. Mechanisms un- derlying the early phase of spike frequency adaptation in mouse spinal motoneurones. J Physiol 566: 519 -- 532, 2005. Mitchell SJ, Silver RA. Shunting inhibition modulates neuronal gain during synaptic excitation. Neuron 38: 433 -- 445, 2003. Naud R, Gerstner W. Coding and decoding with adapting neurons: a population approach to the peri- stimulus time histogram. PLoS Comput Biol 8: e1002711, 2012. Naud R, Marcille N, Clopath C, Gerstner W. Firing patterns in the adaptive exponential integrate-and-fire model. Biol Cybern 99: 335 -- 347, 2008. 12 Ostojic S. Inter-spike interval distributions of spiking neurons driven by fluctuating inputs. J Neurophysiol 106: 361 -- 373, 2011. Schwindt PC, Spain WJ, Crill WE. Calcium- dependent potassium currents in neurons from cat sen- sorimotor cortex. J Neurophysiol 67: 216 -- 226, 1992. Shadlen MN, Newsome WT. The variable discharge of cortical neurons: implications for connectivity, com- putation, and information coding. J Neurosci 18: 3870 -- 3896, 1998. Smith MR, Nelson AB, Du Lac S. Regulation of firing response gain by calcium-dependent mechanisms in vestibular nucleus neurons. J Neurophysiol 87: 2031 -- 2042, 2002. Soma S, Shimegi S, Osaki H, Sato H. Cholinergic modulation of response gain in the primary visual cortex of the macaque. J Neurophysiol 107: 283 -- 291, 2012. Sripati AP, Johnson KO. Dynamic gain changes during attentional modulation. Neural Comput 18: 1847 -- 1867, 2006. Stocker M. Ca(2+)-activated K+ channels: molecular determinants and function of the SK family. Nat Rev Neurosci 5: 758 -- 770, 2004. Sutherland C, Doiron B, Longtin A. Feedback- induced gain control in stochastic spiking networks. Biol Cybern 100: 475 -- 489, 2009. Touboul J, Brette R. Dynamics and bifurcations of the adaptive exponential integrate-and-fire model. Biol Cybern 99: 319 -- 334, 2008. Tuckwell HC. Introduction to Theoretical Neurobiology, volume 2. Cambridge, UK: Cambridge University Press, 1988. Urdapilleta E. Survival probability and first-passage- time statistics of a Wiener process driven by an expo- nential time-dependent drift. Phys Rev E 83: 021102, 2011. Wang XJ. Calcium coding and adaptive temporal com- putation in cortical pyramidal neurons. J Neurophysiol 79: 1549 -- 1566, 1998. Wang XJ, Liu Y, Sanchez-Vives MV, McCormick DA. Adaptation and temporal decorrelation by single neurons in the primary visual cortex. J Neurophysiol 89: 3279 -- 3293, 2003. Wilson NR, Runyan CA, Wang FL, Sur M. Division and subtraction by distinct cortical inhibitory networks in vivo. Nature 488: 343 -- 348, 2012. Ostojic S, Brunel N. From spiking neuron mod- els to linear-nonlinear models. PLoS Comput Biol 7: e1001056, 2011. Pillow JW, Shlens J, Paninski L, Sher A, Litke AM, Chichilnisky EJ, Simoncelli EP. Spatio- temporal correlations and visual signalling in a complete neuronal population. Nature 454: 995 -- 999, 2008. Pospischil M, Piwkowska Z, Bal T, Destexhe A. Comparison of different neuron models to conductance- based post-stimulus time histograms obtained in corti- cal pyramidal cells using dynamic-clamp in vitro. Biol Cybern 105: 167 -- 180, 2011. Prescott SA, Sejnowski TJ. Spike-rate coding and spike-time coding are affected oppositely by different adaptation mechanisms. J Neurosci 28: 13649 -- 13661, 2008. Renart A, Brunel N, Wang XJ. Mean-field theory of irregularly spiking neuronal populations and work- ing memory in recurrent cortical networks. In: Com- putational Neuroscience - A Comprehensive Approach, edited by J Feng, New York, USA: CRC Press, 425 -- 484, 2004. Reynolds JH, Heeger DJ. The normalization model of attention. Neuron 61: 168 -- 185, 2009. Richardson M. Dynamics of populations and networks of neurons with voltage-activated and calcium-activated currents. Phys Rev E 80: 1 -- 16, 2009. Risken H. The Fokker-Planck Equation. New York, USA: Springer, 1996. Roxin A, Brunel N, Hansel D, Mongillo G, van Vreeswijk C. On the distribution of firing rates in networks of cortical neurons. J Neurosci 31: 16217 -- 16226, 2011. Sanchez-Vives MV, McCormick DA. Cellular and network mechanisms of rhythmic recurrent activity in neocortex. Nat Neurosci 3: 1027 -- 1034, 2000. Sanchez-Vives MV, Nowak LG, McCormick DA. Membrane mechanisms underlying contrast adaptation in cat area 17 in vivo. J Neurosci 20: 4267 -- 4285, 2000. Schwalger T, Fisch K, Benda J, Lindner B. How noisy adaptation of neurons shapes interspike interval histograms and correlations. PLoS Comput Biol 6: e1001026, 2010. Schwartz O, Pillow JW, Rust NC, Simoncelli EP. Spike-triggered neural characterization. J Vis 6: 484 -- 507, 2006. Schwindt PC, Spain WJ, Crill WE. Long-lasting reduction of excitability by a sodium-dependent potas- sium current in cat neocortical neurons. J Neurophysiol 61: 233 -- 244, 1989. 13 Figure 1: Steady-state spike rates and ISI distributions of single neurons. A, from top to bottom: Spike times, instantaneous spike rate (r∆t) histogram, membrane voltage (Vi), membrane voltage histogram and adaptation current (wi) of an (adapted) aEIF neuron with a = 0.06 mS/cm2, b = 0 (left) and a = 0, b = 0.18 µA/cm2 (right) √ ms for N = 5000 trials. Spike times and driven by a fluctuating input current with µ = 2.5 mV/ms, σ = 2 mV/ adaptation current are shown for a subset of 10 trials, the membrane voltage is shown for one trial. Results from numerical simulations are shown in grey. Results obtained using the Fokker-Planck equation are indicated by orange lines and include the instantaneous spike rate (r), the membrane potential distribution (p) and the mean adaptation current ( ¯w). r, p and ¯w were calculated from the eqs. (13), (8) and (10), respectively. These quantities have reached their steady state here. The time bin for r∆t was ∆t = 2 ms, for the other parameter values see Materials and Methods. B, top panel: ISI histogram corresponding to the N trials in A and ISI distribution (pISI, orange line) calculated via the first passage time problem (eq. (21)). B, center and bottom panels: Membrane voltage and adaptation current trajectories from one trial in A, but rearranged such that just after each spike the time is set to zero. Histograms for the adaptation current just after the spike times are included. The time-varying mean adaptation current from the first passage time problem (eq. (20)) and the steady-state mean adaptation current from A (eq. (10)) are indicated by solid and dashed orange lines, respectively. All histograms (in A and B) represent the data from all N trials. 14 Figure 2: Spike rate adaptation, gain and threshold modulation in single neurons. A: Cartoon of a single neuron visualizing the input parameters and output quantities. B: Instantaneous spike rate r (top panel), mean membrane voltage (cid:104)V (cid:105)p (center panel, squares) and mean adaptation current ¯w (center panel, solid lines) of an aEIF neuron without adaptation, a = b = 0 (left), and with either a purely subthreshold adaptation current, a = 0.06 mS/cm2, b = 0 (center) or a spike-triggered adaptation current, a = 0, b = 0.3 µA/cm2 (right), in response to a sudden increase in synaptic drive (bottom panel). C: I-O curves of the neurons in B, i.e., spike rate r as a function of presynaptic spike rates rE , rI. Here, rE = rI, but excitation is stronger than inhibition, due to the coupling parameter values (see Materials and Methods). The I-O curves represent the spike rate response of the neurons to a sudden increase of rE and rI, measured in steps of 50 ms after that increase (light to dark colors). Dots indicate the evolution of the spike rate corresponding to the input in B. D: Steady-state spike rate r∞ as a function of the mean µ and standard deviation σ of the fluctuating input. Note that µ and σ are determined by the number of presynaptic neurons, their (Poisson) spike rates and synaptic strengths, cf. eqs. (5) -- (6). The dashed lines in D indicate the values of µ and σ which correspond to the presynaptic spike rates in C, circles mark the values of the moments corresponding to the increased input in B. 15 Figure 3: Changes of spiking variability in single neurons. A: ISI distribution (pISI) of a single aEIF neuron in response to a fluctuating input with mean µ = 0.75 mV/ms and standard deviation σ = 3.25 mV/ ms, for a = 0, 0.03, 0.06 mS/cm2, b = 0 (top) and a = 0, b = 0, 0.15, 0.3 µA/cm2 (bottom). B: ISI coefficient of variation (CV) as a function of µ and σ, for a neuron without adaptation, a = b = 0 (left), and with either a subthreshold adaptation current, a = 0.06 mS/cm2, b = 0 (center) or a spike-triggered adaptation current, a = 0, b = 0.3 µA/cm2 (right). Circles indicate the values of µ and σ used in A. C: Change of ISI CV caused by a subthreshold (left) or spike-triggered (right) adaptation current as a function of µ and σ. The white regions in B and C indicate the parameter values for which the ISI CV was not computed, because r∞ < 1 Hz. √ 16 Figure 4: Gain and threshold modulation caused by network interaction. A: Cartoon of the network visual- izing the coupling parameters. B, top panel: Steady-state spike rate of excitatory aEIF neurons, rpopE,∞ (solid lines) and inhibitory aEIF neurons, rpopI,∞ (dashed lines), as a function of rextEE , for rextIE = 6, 10, 14 Hz (left), J recIE = 0.05, 0.1, 0.2 mV (center), J recEI = −0.45,−0.6,−0.75 mV (right). Inset cartoons visualize the varied parameters as specified on the top left. If not indicated otherwise, J recEI = −0.6 mV, rextIE = 10 Hz and J recIE = 0.1 mV. For the other parameter values see Materials and Methods. B, bottom panel: Steady-state spike rate rpopE,∞ as a function of the input parameters µ and σ for the excitatory neurons. Solid lines and dots in the top panel correspond to those of equal color in the bottom panel. 17 Figure 5: Changes of spiking variability caused by network interaction. A: ISI distributions (pISI) of excitatory aEIF neurons for rextEE = 50 Hz. J recEI = −0.6 mV, rextIE = 10 Hz and J recIE = 0.1 mV if not indicated otherwise. B: ISI CV for excitatory neurons (CVpopE ) as a function of rextEE . Color code as in A. Dots indicate the input and ISI CV values for the ISI distributions in A. Insets: ISI CV as a function of the input parameters µ and σ for the excitatory neurons. Lines and dots (insets) correspond to those of equal color in B. Figure 6: Subthreshold and spike-triggered components of IKCa, IKNa and IM. A: Conductances for the slow K+-currents INa, IKCa, IKNa and IM in steady state as a function of the membrane voltage, normalized to a peak value of 1 mS/cm2. B and C: Subthreshold conductance a and spike-triggered absolute increment ∆IKs (B) and relative increment ∆I rel Ks (C) obtained from the fitting procedure (see Materials and Methods) for the conductance-based model neurons with ¯gKCa ∈ [2, 8] mS/cm2 and ¯gKNa = ¯gM = 0 (dots), ¯gKNa ∈ [2, 8] mS/cm2 and ¯gKCa = ¯gM = 0 (squares), ¯gM ∈ [0.1, 0.4] mS/cm2 and ¯gKCa = ¯gKNa = 0 (diamonds). Darker symbols indicate larger conductance values. 18 Figure 7: Subthreshold and spike-triggered components of a range of slow K+-currents A: Steady-state K+-conductance gKs,∞(V ) = ¯gKsω∞(V ) as a function of the membrane voltage, for the generic Hodgkin-Huxley-type description of a slow K+-current (see Materials and Methods), with half-activation voltage α = −40 mV (left curves), α = −10 mV (right curves), inverse steepness β = 6, 9, 12 mV and peak conductance ¯gKs = 1 mS/cm2. The dashed curve indicates the Na+-conductance gNa,∞(V ) of the conductance-based model, normalized to a maximum value of 1 mS/cm2. B: Subthreshold conductance a obtained from the fitting procedure for different values of the parameters α and β. C: Absolute and relative spike-triggered increments ∆IKs (top panel) and ∆I rel Ks (bottom panel), respectively, as a function of α, for τω = 100 ms (left) and τω = 300 ms (right). 19 Figure 8: Effects of IKCa, IKNa and IM on I-O curve and ISI variability. A: Spike rate of a conductance-based model neuron without slow K+-currents, ¯gKCa = ¯gKNa = ¯gM = 0 (black), and with either type of slow K+-current included, ¯gKCa = 8 mS/cm2 (red), ¯gKNa = 8 mS/cm2 (blue), ¯gM = 0.4 mS/cm2 (green), in response to a sudden √ increase of mean input µ, measured in four subsequent time intervals of 250 ms after that increase (light to dark colors). The baseline mean input was µ = 0.05 mV/ms and the input standard deviation was σ = 0.5 mV/ ms. Average values over 50 independent trials are shown. The adapting I-O curve of the neuron with increased IKNa (¯gKNa = 8 mS/cm2) converges very slowly to the steady-state curve (dashed blue) measured 20 s after the increase in √ µ. B: ISI CV of the neurons in A as a function of mean input µ for low (left), medium (center), and high (right) noise ms), respectively. The ISIs were collected over an interval of 10 s after the steady-state intensity (σ = 1, 1.5, 2 mV/ spike rates were reached, in 50 independent trials. 20
1811.08498
1
1811
2018-11-20T21:44:16
Global Sensitivity Analysis of High Dimensional Neuroscience Models: An Example of Neurovascular Coupling
[ "q-bio.NC" ]
The complexity and size of state-of-the-art cell models have significantly increased in part due to the requirement that these models possess complex cellular functions which are thought--but not necessarily proven--to be important. Modern cell models often involve hundreds of parameters; the values of these parameters come, more often than not, from animal experiments whose relationship to the human physiology is weak with very little information on the errors in these measurements. The concomitant uncertainties in parameter values result in uncertainties in the model outputs or Quantities of Interest (QoIs). Global Sensitivity Analysis (GSA) aims at apportioning to individual parameters (or sets of parameters) their relative contribution to output uncertainty thereby introducing a measure of influence or importance of said parameters. New GSA approaches are required to deal with increased model size and complexity; a three stage methodology consisting of screening (dimension reduction), surrogate modeling, and computing Sobol' indices, is presented. The methodology is used to analyze a physiologically validated numerical model of neurovascular coupling which possess 160 uncertain parameters. The sensitivity analysis investigates three quantities of interest (QoIs), the average value of $K^+$ in the extracellular space, the average volumetric flow rate through the perfusing vessel, and the minimum value of the actin/myosin complex in the smooth muscle cell. GSA provides a measure of the influence of each parameter, for each of the three QoIs, giving insight into areas of possible physiological dysfunction and areas of further investigation.
q-bio.NC
q-bio
Global Sensitivity Analysis of High Dimensional Neuroscience Models: An Example of Neurovascular Coupling J.L. Hart · P.A. Gremaud · T. David 8 1 0 2 v o N 0 2 ] . C N o i b - q [ 1 v 8 9 4 8 0 . 1 1 8 1 : v i X r a Abstract The complexity and size of state-of-the-art cell models have significantly increased in part due to the requirement that these models possess complex cellular functions which are thought -- but not necessarily proven -- to be important. Modern cell models often involve hundreds of parameters; the values of these parameters come, more often than not, from animal experiments whose relationship to the human physiology is weak with very little information on the errors in these measurements. The concomitant uncertainties in parameter values result in uncertainties in the model outputs or Quantities of Interest (QoIs). Global Sensitivity Analysis (GSA) aims at apportioning to individual parameters (or sets of parameters) their relative contribution to output uncertainty thereby introducing a measure of influence or importance of said parameters. New GSA approaches are required to deal with increased model size and complexity; a three stage methodology consisting of screening (dimension reduction), surrogate modeling, and computing Sobol' indices, is presented. The methodology is used to analyze a physiologically validated numerical model of neurovascular coupling which possess 160 uncertain parameters. The sensitivity analysis investigates three quantities of interest (QoIs), the average value of K + in the extracellular space, the average volumetric flow rate through the perfusing vessel, and the minimum value of the J.L.H. was supported in part by the National Science Foundation (NSF) award NSF DMS-1522765. P.A.G. was supported in part by National Science Foundation (NSF) awards NSF DMS-1522765 and DMS-1745654. J.L. Hart Department of Mathematics, North Carolina State University, Raleigh, NC, USA, E-mail: [email protected], P.A. Gremaud Department of Mathematics, North Carolina State University, Raleigh, NC, USA, E-mail: [email protected] T. David Department of Mechanical Engineering, University of Canterbury, New Zealand, E-mail: [email protected] actin/myosin complex in the smooth muscle cell. GSA provides a measure of the influence of each parameter, for each of the three QoIs, giving insight into areas of possible physiological dysfunction and areas of further investigation. Keywords: neurovascular coupling, global sensitivity analysis, model parameters 1 Introduction Over the past 20 years, the use of computational models to describe physiological phenomena has grown spectacularly. This growth has provided significant advantages to the experimental community in that it can furnish results of in silico experiments which are either ethically or physically impossible in the laboratory. However, these models have also increased in complexity with a concomitant increase in the associated number of parameters (for a variety of reasons, most notably the requirement that the models should possess complex cellular functions which are thought but not necessarily proven to be important). These parameters, in defining the relevant phenomenon, more often than not come from a plethora of animal experiments whose relationship to the human physiology is weak. In addition, experimental results in the public domain provide very little information on the errors in these measurements. Even simple physiological models tend to be highly non-linear; their range of applicability and reliabil- ity can only be assessed through careful analysis. For large complex systems, the sensitivity of quantities of interest to model parameters is a priori unclear. Herein lies one of the difficulties of modeling: what effect do uncertainties in parameters determined from experiment have on the output of a non-linear numerical model? Because of both model complexity and high dimensionality, sensitivity analysis is com- putationally demanding and may require several ad hoc steps -- such as screening (reducing the parameter dimension). Importantly, in analysing sensitivities, we can learn significant facts about the physiology of the system which would have stayed hidden under the premise of simply producing results. From a purely physiological perspective, an understanding of the dominant cellular mechanisms resulting in cerebral tissue perfusion after neuronal stimulation would be of particular and important interest. In the above context, we investigate the sensitivity of the neurovascular coupling (NVC) response (see Section 2), which has a large parameter dimension where most (if not all) of the parameter values 2 come from non-human experiment with an inherent (unknown) error. We denote by y = (y1, . . . , yN ) the state variables of the model and by θ = (θ1, . . . , θP ) the uncertain parameters of the model. The evolution of the state variables is governed by a system of ordinary differential equations (ODEs) dy dt = f (y, θ), (1.1) where f is a known function of its arguments. Equation (1.1) is completed with a set of initial conditions y(0) = y0; here, we simply take y0 as the equilibrium solution at the parameters' nominal values, i.e., f (y0, ¯θ) = 0, where ¯θ = (¯θ1, . . . , ¯θP ) denotes the nominal values of the parameters. The Supplementary Material contains code and the nominal values of parameters, further information can be found in [6]. Based on physiological considerations, we examine three quantities of interest (QoI), see Section 2. Let q be one of our three QoIs; while determined from the state variables, i.e., from y, q is ultimately a function of the parameters alone (and possibly time), i.e. q = g(θ). (1.2) The overall goal of our numerical study is to determine which of the uncertain parameters θ1, . . . , θP are the most/least influential for each QoI. There is a substantial amount of research currently being done in applied mathematics and statistics in the corresponding field of Global Sensitivity Analysis (GSA). How to meaningfully define "influential" or "non-influential" and how to develop methods applicable to high-dimensional problems are two significant challenges of the field [1, 14, 15, 17, 32, 35]. The present NVC model contains N = 67 state variables and P = 160 uncertain parameters. The complexity of the model and large parameter space dimension preclude a direct application of GSA tools. Part of our contribution in this paper is to show how multiple GSA tools may be combined to analyse such problems. 2 Physiological Model: Neurovascular Coupling The NVC response, i.e., the ability to locally adjust vascular resistance as a function of neuronal activ- ity, is believed to be mediated by a number of different signaling mechanisms. A mechanism based on a metabolic negative feedback theory was first proposed in [34]. According to this theory, neural activity leads to a drop in oxygen or glucose levels and increases in CO2, adenosine, and lactate levels. All of these 3 signals could dilate arterioles and hence were believed to be part of the neurovascular response. How- ever, recent experiments illustrated that the NVC response is partially independent of these metabolic signals [22, 23, 25, 31, 33]. An alternative to this theory was proposed where the neuron releases signal- ing molecules to directly or indirectly affect the blood flow. Many mechanisms such as the potassium (K+) signaling mechanism [11], the nitric oxide (NO) signaling mechanism, or the arachidonic acid to epoxyeicosatrienoic acid (EET) pathway are found to contribute to the neurovascular response [3]. The K+ signaling mechanism of NVC seems to be supported by significant evidence, although new evidence shows that the endfoot astrocytic calcium (Ca2+) could play a significant role. The K+ signaling hypothesis mainly utilises the astrocyte, positioned to enable the communication between the neurons and the local perfusing blood vessels. The astrocyte and the endothelial cells (ECs) surrounding the perfusing vessel lumen exhibit a striking similarity in ion channel expression and thus can enable control of the smooth muscle cell (SMC) from both the neuronal and blood vessel components [24]. Whenever there is neuronal activation K+ ions are released into the extracellular space (ECS) and synaptic cleft (SC). The astrocyte is depolarised by taking up K+ released by the neuron and releases it into the perivascular space (PVS) via the endfeet through the BK channels [12]. This increase in ECS K+ concentration (3− 10 mM) near the arteriole hyperpolarises the SMC through the inward rectifying K+ (KIR) channel, effectively closing the voltage-gated Ca2+ channel, reducing smooth muscle cytosolic Ca2+ and thereby causing dilation. Higher K+ concentrations in the PVS cause contraction due to the reverse flux of the KIR channel [10]. In maintaining a relatively homeostatic condition the neuron uses a considerable amount of energy ( as noted below) on ensuring a specific concentration of ions (and ion gradients) in the extracellular space. This concentration of K+and Na+can be envisaged as one of the inputs to the astrocyte in determining the nutrient flux to the cerebral tissue. Parameters which model/affect the energy usage are therefore important in simulating neurovascular coupling. In order not to increase the complexity of the numerical model further than is necessary an approximation needs to be made on how to model the energy usage and the associated parameters. Estimates of the relative demands of the cerebral processes that require energy were given based on different experimental data by Ames [2]. Vegetative processes that maintain this homeostasis including protein synthesis accounted for 10 − 15% of the total energy consumption. 4 The costliest function seems to be in restoring the ionic gradients during neural activation. The sodium potassium (Na+/K+) exchange pump is estimated to consume 40 − 50%, while the Ca2+ influx from organelles and extracellular fluid consumes 3 − 7%. Processing of neurotransmitters such as uptake or synthesis consumes 10 − 20%, while the intracellular signaling systems which includes activation and inactivation of proteins consumes 20 − 30%. The rest of the energy is estimated to be consumed by the axonal and dendritic transport in both directions. Given the distribution of energy amongst the cellular pathways it is assumed for this model that variation in oxygen concentration in the cerebral tissue affects only the Na+/K+ exchange pump. Previous work [30] has provided the construction of an experimentally validated numerical (in silico) model based on experimental data to simulate the functional magnetic resonance imaging (fMRI) blood- oxygen-level dependent (BOLD) signal associated with NVC along with the associated metabolic and blood volume responses. An existing neuron model [28, 29] has been extended to include an additional transient sodium (Na+) ion channel (NaT) expressed in the neuron, and integrated into a complex NVC model [7, 8, 19]. This present model is based on the hypothesis that the K+ signaling mechanism of NVC is the primary contributor to the vascular response and the Na+/K+ exchange pump in the neuron is the primary consumer of oxygen during neural activation. The model contains 317 parameters, most of which come from non-human experiments. Based on the work by [8] and [20], we have chosen a subset of parameters defining basic pathways, such as the nitric oxide and potassium pathways, that are considered important for the normal function of neurovascular coupling. We model the uncertainty of the chosen parameters by representing them as random variables. The remaining parameters are fixed to nominal values; they include leak terms, characteristic oxygen and other species concentrations, buffer concentrations, volume surface ratios etc. . . By permitting variability in only the parameters that support these pathways, the dimension of the parameter space is reduced from 317 to 160, which greatly facilitates our analysis. The algorithms defined below can be used to investigate other complex models including that of neurovascular coupling. However, for this initial work, we constrain ourselves to the above subset. Figure 1 shows the components and main pathways of the neurovascular coupling model (version 2.0). The numerical model outlined in sketch form in Figure 1 is fully defined in the Supplementary Material and has been developed over a number of years [7, 8, 10]. 5 Fig. 1: Graphic Sketch of NVU version 2.0, showing the basic components of neuron, astrocyte, smooth muscle cell, endothelial cell, lumen and extracellular space. ion channels, pumps and pathways from neuron to endothelial cell are also shown in addition to the basis for evaluation of the fMRI blood oxygenation level signal (BOLD) protocol. In order to induce a variation in radius of a vessel which perfuses the associated cerebral tissue following a neuronal stimulation, an input current is used. The numerical experiments solve equation (1.1) in the presence of short duration electrical (current) stimuli as displayed in Figure 2. For the presented cases, two input profiles are utilised. A rectangular pulse of width 10 seconds and a experimental pulse sequence (stimulating the whisker pad of a rat) used in the work of [40] which has the same magnitude as the rectangular pulse but a duration of sixteen seconds followed by second pulse (which is not used in this analysis). Further information about the experiment and the results can be found in [40]. For this second case, the stimulus was a current injection at the rat whisker pad. This induced an increase in 6 ERERPumpPumpIPR3LeakSRPumpIP R3LeakCICRCICR0CationExtrusionStretchKResidualCa2+cplVcplIP3cplPLCLeakwall shear stresseNOSNOPVSSMCLUNOECStretchNaCaClk1k-1k3k2k4EEEb6c5cNONONONOBKVOCCNaKExtrusionIP R3KIRTRPV4BKA + MA + MpAMpAMKKKKKKK1425637RadiusEETACSCNONa3HCONBCKCC1NKCC1NaKKClρO2CBFCBVCMRO2HbRBOLDO2O2ECSNMDAPostsynaptic terminalnNOSNODendriteSoma/axonNa leakK leakNaPKAKDRNMDAKDRNaPKAK leakNa leakNaT[KB][B]+[K ]+BufferingInput current I(t)NaK pumpO2O2O2NaK pumpO2O2O2O2Ca2+ionK+ionIP3Na+ionagonistClion'voltage'degradationionchannelionpumpreceptorHCO3ion--GluEETstretchcGMPCaMdiffusionNOO2Nitric oxideOxygenO2O2Glu neuronal activity in the somato-sensory cortex which subsequently, through the neurovascular pathway, produced a change in the radius R(t) allowing increased nutrients to perfuse into the cerebral tissue; this is the essence of neuro-vascular coupling. Fig. 2: Left: rectangular pulse input stimulus; right: stimulus used in lab experiments. We wish to analyse parameter sensitivity of quantities of interest (QoIs) that are key to our un- derstanding and quantification of neurovascular coupling. From previous work, it is known that the K+concentration in the extracellular space (ECS) is crucial to maintaining a homeostatic state (large concentrations of K+support the propagation of spreading depression waves) hence we choose to look at the average value of K+in the ECS as our first QoI q1. Secondly, neurovascular coupling is the main phe- nomenon providing oxygen and nutrients to the neuronal tissue. This is mediated by the local arteriole dilating and (under the assumption of constant pressure) increasing the flow of blood into the tissue. We therefore define the time averaged cerebral blood flow determined over the course of neuronal stimulation as our second QoI, q2. The contraction/dilation of the arteriolar vessel depends on the smooth muscle cell (SMC) concentration of Ca2+and the phosphorylation of the actin/myocin complex. Our third QoI q3 is the minimum combined concentration of the actin myosin complex, both phosphorylated and unphos- phorylated. This allows us to analyse the functioning of the main components of the neurovascular unit (NVU) defined as the linked components of neuron, synaptic ceft, astrocyte, perivascular space (PVS), SMC, endothelial cell (EC), lumen (LU, the domain in which blood flows) and the ECS. Assuming the stimulation occurs for time t between t1 and t2, t2 > t1, the QoIs are thus 7 95100105110115120Time (sec)00.0050.010.0150.020.0250.03Input Stimulus95100105110115120Time (sec)00.0050.010.0150.020.0250.03Input Stimulus -- average ECS potassium q1 = 1 t2 − t1 [K +]ECS(s) ds, (cid:90) t2 t1 (cid:90) t2 -- average volumetric flow rate in the cerebral tissue (cid:18) R(s) (cid:19)4 q2 = 1 t2 − t1 t1 R0 ds, (2.1) (2.2) -- minimum combined concentration of the actin myosin complex, both phosphorylated and unphos- phorylated q3 = min t,t1≤t≤t2 [AM (t) + AMp(t)], (2.3) which we will refer to as [AM + AMp]min throughout the article. Other examples of sensitivity analysis studies relevant to the general field of biomedicine include [16, 37, 38, 39]. 3 Methodology Our analysis may be described in a four step process: (i) define a probability distribution for θ which models its uncertainty, (ii) draw samples from this distribution, (iii) evaluate the QoI for each of these parameter samples, (iv) use these QoI evaluations to infer its sensitivity to each parameter. Subsection 3.1 describes our approach for step (i). Step (ii) is easily executed using standard methods. The computational bottleneck for our analysis is the QoI evaluations (which require ODE solves) in step (iii). While relatively large and stiff, the ODE system (1.1) can be solved with standard tools and methods, here through the MATLAB routine ode15s with relative and absolute tolerances of 10−4. The evaluation of the above QoIs themselves from the ODE solutions is straightforward and can be done at low cost. Step (iv), inferring the sensitivities, may be done in a plurality of ways. Because of the parameter dimension and computational cost of model evaluations, we perform step (iv) with a multi-phased procedure, (I) screening, 8 (II) surrogate modeling, (III) and computing Sobol' indices, described in Subsections 3.2, 3.3, and 3.4 respectively. We summarize our method in Subsection 3.5. 3.1 Parameter Distribution Fitting Describing the uncertainty attached to the parameter vector θ is an important and delicate modeling assumption. Here, we give each θi, i = 1, . . . , 160, a nominal value ¯θi and assume each parameter to be independent and uniformly distributed over the interval [0.9 ¯θi, 1.1 ¯θi], i.e., within ±10% of the nominal value. Larger intervals were considered for the parameters. For the NVU model under consideration, increasing the width of the interval resulted in many samples for which the solution exhibited atypical or non-physical behavior, or in some cases ode15s was unable to solve the system. Our choice of ±10% uncertainty is considered a reasonable compromise between accounting for uncertainty and ensuring computational feasibility. Our initial assumption of parameter independence is incorrect. In fact, ode15s is unable to solve the system for most samples drawn under this assumption. The parameter dependencies, which are unknown a priori, are discovered through a computational procedure akin to Approximate Bayesian Computation [27]. Our approach, described below, is computationally intractable when applied to the NVU model under consideration. We simplify the model by removing the stimulus, i.e. modeling the steady state behavior of the system. A collection of S samples, θk, k = 1, 2, . . . , S, are drawn from our initial distribution (assuming independence) and the ODE system is solved for each parameter sample. The collection of solutions are post-processed to partition the samples {θk}S k=1 into two subsets {θak}Sa and {θrk}Sr k=1, where ak, k = 1, 2, . . . , Sa denotes the samples where the steady state solution exhibited k=1 physiologically normal behavior and rk, k = 1, 2, . . . , Sr denotes the samples where it does not. We reject the samples {θrk}Sr k=1 and fit a distribution to the accepted samples {θak}Sa k=1 using standard statistical methods. This new distribution is sampled S times, the ODE system is solved for each sample, and the results are post-processed into accepted and rejected samples again. We continue this process iteratively until satisfactory convergence of the fitted distribution. 9 160(cid:88) 3.2 Screening Having determined a parameter distribution, we evaluate the QoI q = g(θ) at M different samples, denote them θk, k = 1, 2, . . . , M . We fit a linear model to the QoI under study g(θk) = g(θk 1 , . . . , θk P ) ≈ β0 + βjθk j , k = 1, . . . , M. (3.1) j=1 This approach yields a crude (but highly efficient here) sensitivity analysis of the model with respect to the θj's, j = 1, . . . , 160. We assign a preliminary importance measure to each θj by computing for each of them the relative size of their coefficient in the above linear approximation, i.e. Lj = 160(cid:80) βj β(cid:96) , (cid:96)=1 j = 1, . . . , 160. To obtain a model with a more manageable size, we reduce the parameter space to only the θj's for which Lj > 0.01. We denote these r parameters {θji}r i=1. The rest of the parameters are regarded as non-influential and treated as latent variables, even though they are uncertain, their specific values (within the given range) have little bearing of the considered QoI. In other words, we consider the approximation g(θ1, . . . , θ160) ≈ h(θj1, . . . , θjr ), (3.2) where h is obtained from g by treating the non-influential parameters as latent. In Section 4, this reduction yields around 15-20 parameters instead of the original 160. We use θ to denote the reduced parameter vector. 3.3 Surrogate model For any of the three considered QoIs, our information on the function h defined (3.2) consists of the set of sampled values {h(θk )}, k = 1, . . . , M . To facilitate the use of standard GSA tools, which , . . . , θk jr j1 may require derivatives or variance estimations, it is both convenient and computationally advantageous to construct an approximating function, i.e., a surrogate model. We use a sparse Polynomial Chaos (PC) surrogate. This amounts to introducing a polynomial approximation of h of the type h( θ) ≈ H( θ) ≡(cid:88) cαψα( θ) (3.3) α 10 where the ψα's are multivariate polynomials which are orthogonal with respect to the probability distri- bution function (PDF) p θ of θ, i.e. (cid:90) ψα(x)ψβ(x) p θ(x) dx = δα,β (3.4) where α and β are multi-indices and δα,β is the generalized Kronecker symbol. The coefficients are computed through least-squares minimization, see the Appendix for additional discussion. All surrogate models are validated using 10-fold cross validation. Polynomial Chaos is by now a well documented method. We use the UQLab implementation for the results below and refer the reader to its manual [26] for more details. 3.4 Sobol' indices We use variance based GSA to assess the relative importance of the input parameters of H in (3.3). In their simplest form, the total Sobol' indices [36] apportion to uncertain parameters, or sets thereof, their relative contribution to the variance of the output. Indeed, thanks to the law of total variance, we can decompose the variance of H( θ) as var(H( θ)) = var(E[H( θ) θ∼i]) + E[var(H( θ) θ∼i)], (3.5) where θ∼i denotes all the parameters in θ except θi. If we assume now that all the input parameters of H are known with certainty, i.e., if θ∼i is known, then the remaining variance of H( θ) is simply given by var(H( θ)) − var(E[H( θ) θ∼i]) = E[var(H( θ) θ∼i)]. This latter expression is thus a natural way of measuring how influential θi is; the corresponding total Sobol' index STi is but a normalized version of this STi = E[var(H( θ) θ∼i)] var(H( θ)) . (3.6) From this definition, one easily observes that STi ∈ [0, 1]; large values indicate that θi is important, STi ≈ 0 implies that θi is not important. The relevance of this basic definition can be extended to time dependent QoIs [1], stochastic QoIs [15] or correlated parameters [14]. 11 3.5 Summary of the Method Algorithm 1 provides a summary of the method. Algorithm 1 overall numerical approach 1: while parameter distribution has not converged do 2: 3: 4: 5: 6: for k = 1 : S do (cid:46) sampling sample parameter distribution −→ θk; solve (1.1) (without stimulus) −→ yk end for partition the parameter samples into accepted and rejected samples fit a new distribution to the accepted samples (cid:46) see § 3.1 7: end while 8: for k = 1 : M do 9: sample parameter distribution −→ θk; solve (1.1) (with stimulus) −→ yk (cid:46) sampling (final distribution) 10: end for 11: for each QoI q do 12: 13: 14: 15: solve the least-squares problem (3.1) identify the influential parameter vector θ = (θj1 , . . . , θjr ) fit the polynomial chaos surrogate H (3.3) compute total Sobol' indices (3.6) of the surrogate model H 16: end for (cid:46) linear model (cid:46) screening (cid:46) surrogate model (cid:46) Sobol' indices 4 Numerical results 4.1 Parameter distribution Applying the methodology described in Section 3.1, S = 919 parameter samples are drawn and (1.1) is solved for each sample with no stimulus applied. Of these 919 runs, 670 of them fail due to ode15s terminating prematurely because its time steps became too small. The remaining 249 yields 139 solutions for which the radius reaches a stable steady state (at least for the duration of the time integration) and 110 solutions for which the radius reaches a steady state but subsequently becomes transient. The left panel of Figure 3 displays four representative solutions for the radius; the red curves remain in steady state for the duration of our time integration whereas the black curves revert into a transit regime after 12 some time in or near steady state. The right panel of Figure 3 displays the samples for two parameters (defined in the buffer equation (4.1), see the Supplementary Material for details) which are found to be highly influential in determining the behavior of the solution. A blue * indicates a sample where the solver terminated prematurely, a yellow ♦ indicates a sample where an unstable steady state was observed, a red ◦ indicates a sample where the solution remained in steady state. A strong correlation determining the behavior of the solution is observed. The two parameters in question, denoted by (θ120, θ121), occur in the buffer equation defined by ff = 1+exp( θ118 −[K+ ]e+θ120 θ121 ) [K +]e + B (cid:10) [K +]b fr = θ118 d[K +]b dt = ff [K +]e(θ119 − [K +]b) − fr[K +]b The parameters θ120 and θ121 have values 5.5 and 1.09, representing a shift and scaling in the forward rate constant, respectively. (4.1) Fig. 3: Left: examples of stable (red) and unstable (black) steady state solutions. Right: samples of the buffer parameters (θ120, θ121) using uniform independent sampling. A blue * indicates the sample yielded a premature termination of the solver, a yellow ♦ indicates the sample yielded an unstable steady state, a red ◦ indicates the sample yielded a stable steady state. Observing this correlation, we use the accepted samples to fit (θ120, θ121) with a bivariate Frank copula with beta marginals. The experiment is repeated by sampling the two correlated parameters from 13 02004006008001000Time (sec)1.822.22.42.62.83Radius (m)10-555.25.45.65.86p111.051.11.151.2p2 this bivariate distribution and all other parameter from their original uniform distributions. After four iterations refining the joint distribution of (θ120, θ121), we were able to generate 902 out of 960 samples which yielded solutions with stable steady states (51 solutions had unstable steady states and 7 had premature solver terminations). This fitted distribution is used for all subsequent analysis. Samples are drawn and the model, with a stimulus applied (in two separate cases, the 10 second rectangular pulse and the 16 second experimental pulse), is run for each sample. This results in solutions exhibiting three different physiological regimes; they are displayed in Figure 4 where the radius is plotted as a function of time. The leftmost panel corresponds to the typical case when the radius increases in response to the stimulus and then decreases when it is removed; the center panel corresponds to an atypical case where the radius has an initial decrease in response to the stimulus; the right panel corresponds to another atypical case where the radius reaches another steady state and does not decrease after the stimulus is removed. Fig. 4: Radii corresponding to samples (using the rectangular pulse stimulus). Left: curves an an increase in response to the stimulus; center: curves with a decrease in response to the stimulus; right: curves which settle in a different steady state. This article focuses on the non-pathological case, corresponding to the left panel of Figure 4, so we remove samples where the radius does not increase in response to the stimulus and decrease when it is removed. However, we recognise that a decrease in radius upon stimulation does not necessarily mean an incorrect result. Indeed these cases are of particular importance (due to the possibility of the existence of cortical spreading depression [21]) and a topic of future research. This processing yields 660 samples for analysis when the rectangular pulse stimulus is applied and 438 samples when the experimental pulse stimulus is applied. The results presented below use these samples. 14 Exploration of the 660 retained samples indicate that "pathological" cases have higher probability when the parameter which shifts the activation variable for the K+flux through the soma KDR channel is reduced; however, this parameter does not characterize the solution regime by itself; it is likely that the solution regime is characterized by a combination of several parameters. Further sampling and exploration is required to better understand the structure in parameter space which determine the solution regime. The 160 parameters are indexed (purely for coding reasons) and are not to be taken as a ranked order. The tables of ranked parameters, given in the subsections below to summarize the most influential parameters, provide the parameters in the first column, their location in the Supplementary Material in the second column, and their total Sobol' indices for the experimental and rectangular pulses in the third and fourth columns, respectively. Each table contains the five most influential parameters for a given QoI, as measured by the total Sobol' indices for the experimental pulse. There are slight differences in the ordering of the less important parameters for the rectangular pulse and experimental pulse cases; the tables below report the ranking from the experimental pulse case. Both rectangular and experimental pulse cases agree on the ranking of the most influential parameters. 4.2 Average ECS Potassium Figure 5 and 6 display results for the average of the ECS potassium, defined in equation (2.1), for the rectangular pulse stimulus and experimental pulse stimulus, respectively. In the top left panels, predic- tions of the linear surrogate are plotted against the QoI values. The sensitivities Lj, j = 1, 2, . . . , 160, are displayed in the top right panels. Predictions of the PC surrogate are plotted against the QoI values in the bottom left panels. The total Sobol' indices of the PC surrogate are given in the bottom right panels. Table 1 reports the five most important parameters and their respective total Sobol' indices. The first and third parameters in Table 1 are scaling and shift parameters (θ62 and θ63) for the activation gating variable, m4 in the dendrite NaP channel respectively, whose ODE is defined as dm4 dt = m4α(1 − m4) − m4βm4, m4α = m4β = 1 6(1 + exp(−(θ62vd + θ63))) exp(−(θ62vd + θ63)) 6(1 + exp(−(θ62vd + θ63))) , . 15 (4.2) Fig. 5: Average ECS potassium QoI with a rectangular pulse stimulus. From left to right and top to bottom: linear surrogate predictions, linear surrogate importance measure, PC surrogate predictions, total Sobol' indices for PC surrogate. where vd is the dendrite membrane potential. The nominal values of θ62 and θ63 are 0.143 and 5.67, respectively. These effectively define the charac- teristic time scale and forcing function in the rate equation for the open probability of the persistent sodium channel. The second most important parameter determines the strength of the conductance in the K+leak ion channel. The fourth and fifth parameters, θ75 and θ76, are the shift and scale of the neuron membrane potential in the ODE for the activation variable for the K flux through dendritic KDR channel, defined as (cid:18) m6α = θ74 vd + θ75 1 − exp(−(θ76vd + θ76θ75)) (cid:19) , (4.3) where the nominal values for θ74, θ75, and θ76 are 0.016, 34.9, and 0.2, respectively. 16 55.566.577.58Model Value55.566.577.58Surrogate PredictionLinear SurrogateLinear Surrogate120406080100120140160Parameter Index00.020.040.060.080.1Importance Measure55.566.577.58Model Value55.566.577.58Surrogate PredictionPC SurrogatePC Surrogate61626374757677787994116119120141Parameter Index00.10.20.3Total Sobol' Index Fig. 6: Average ECS potassium QoI with experimental pulse stimulus. From left to right and top to bottom: linear surrogate predictions, linear surrogate importance measure, PC surrogate predictions, total Sobol' indices for PC surrogate. The results for this specific QoI are similar for both the rectangular pulse and experimental pulse stimulus. In both cases, the QoI is approximated with reasonable accuracy by a linear surrogate and with higher accuracy by the PC surrogate. The most important parameters are shared in both cases. Notice that parameter θ120, defined in (4.1), appears to be important in the linear surrogate but unimportant in the PC surrogate. This is because it is strongly correlated with parameter θ121 (also defined in defined in (4.1)), see Figure 3, and as a result the coefficient in the linear surrogate may be very large because its effect is offset by the effect of θ121. To decorrelate inputs, the PC surrogate is build with only θ120 instead of both θ120 and θ121. It subsequently has minimal importance. 17 55.566.577.58Model Value55.566.577.58Surrogate PredictionLinear SurrogateLinear Surrogate120406080100120140160Parameter Index00.020.040.060.080.1Importance Measure55.566.577.58Model Value55.566.577.58Surrogate PredictionPC SurrogatePC Surrogate61626374757678849498116117119120122141Parameter Index00.10.20.3Total Sobol' Index Parameter Identification in Supplementary Material Total Sobol' Index (exp.) Total Sobol' Index (rect.) θ62 θ61 θ63 θ75 θ76 Nominal value 0.143 in equation (61) gK,leakd in equation (30) Nominal value 5.67 in equation (61) Nominal value 34.9 in equation (54) Nominal value 0.2 in equation (54) 0.3732 0.2409 0.2297 0.1112 0.0247 0.3738 0.2074 0.2449 0.1124 0.0198 Table 1: Five most influential parameters for the average ECS potassium QoI when the experimental pulse stimulus is applied. The leftmost column is the parameter, the left-center column identifies the parameter in the Supplementary Material, the right-center column is the total Sobol' index computed for the parameter using the experimental pulse stimulus, and the right column is the total Sobol' index computed for the parameter using the rectangular pulse stimulus. We also analysed another QoI for the ECS potassium, namely, its maximum over the interval of stimulation. The results are not reported because of their similarity to the average ECS potassium QoI results given above. The remaining two QoIs also present very similar results for both the rectangular pulse and ex- perimental pulse stimulus. In the interest of conciseness, we only present figures corresponding to the experimental pulse stimulus for these two QoIs; Tables 2 and 3 give results for both stimuli. 4.3 Average Volumetric Flow Rate Figure 7 displays results for the average volumetric flow rate in the cerebral tissue defined by (2.2) in the same manner as Figures 5 and 6. Table 2 reports the five most important parameters and their total Sobol' indices. Unsurprisingly, the parameter list contains values found in the SMC/EC compartment of the full model. However, the topmost parameter, θ141, is associated with the conductance of the inwardly rectifying SMC KIR channel, gKIR, defined as a function of both membrane potential vSM C and the 18 K+concentration in the perivascular space [K +]P V S, given by gKIR = exp(cid:0)θ142vSM C + θ140[K +]P V S − θ141 (cid:1) (4.4) as shown in [7] fitting to the data of [11]. θ141 shifts the conductance to the right for constant [K +]P V S concentration in the perivascular space. The second parameter in Table 2, θ158 (found in the wallme- chanics section of the model) determines the strength of influence of cytosolic [Ca2+] in determining the reaction rate of phosphorylation of myosin [13]. Although not especially important, the third listed parameter θ139 shifts the Nernst potential for the KIR channel to the right in the equation vKIR = θ138[K +]P V S − θ139. (4.5) We again observe reasonably accurate fits by a linear surrogate and improved accuracy by a PC sur- rogate. The surrogate predictions and important parameters for the rectangular pulse and experimental pulse closely agree. As in Subsection 4.2, parameter θ120 appears important in the linear surrogate but unimportant in the PC surrogate. Parameter Identification in Supplementary Material Total Sobol' Index (exp.) Total Sobol' Index (rect.) θ141 θ158 θ139 θ136 θ142 z4 in equation (149) ncross in equation (214) z2 in equation (148) GKi in equation (149) z5 in equation (149) 0.4420 0.3544 0.0436 0.0362 0.0305 0.4561 0.3311 0.0529 0.0305 0.0351 Table 2: Five most influential parameters for the average volumetric flow rate QoI when the experimental pulse stimulus is applied. The leftmost column is the parameter, the left-center column identifies the parameter in the Supplementary Material, the right-center column is the total Sobol' index computed for the parameter using the experimental pulse stimulus, and the right column is the total Sobol' index computed for the parameter using the rectangular pulse stimulus. 19 Fig. 7: Average volumetric flow rate QoI with experimental pulse stimulus. From left to right and top to bottom: linear surrogate predictions, linear surrogate importance measure, PC surrogate predictions, total Sobol' indices for PC surrogate. 4.4 [AM + AMp]min Figure 8 displays results for the minimum of the combined concentration of the actin myosin complex with the experimental pulse stimulus. Table 3 reports the five most important parameters and their total Sobol' indices. By comparing Table 3 and Table 2, we see that the [AM + AMp]min and average volumetric flow rate QoIs share four of their five most important parameters. Although at first analysis this should not be surprising it does suggest that the reaction rates of the actin myosin model for vessel contraction/dilation are relatively insensitive to finding the minimum of the contraction force and that the KIR ion channel is a vital component of the model. 20 1234Model Value11.522.533.54Surrogate PredictionLinear SurrogateLinear Surrogate120406080100120140160Parameter Index00.050.1Importance Measure1234Model Value11.522.533.54Surrogate PredictionPC SurrogatePC Surrogate14376275120135136137138139140141142154156157158160Parameter Index00.10.20.30.4Total Sobol' Index Fig. 8: [AM + AMp]min QoI with experimental pulse stimulus. From left to right and top to bottom: linear surrogate predictions, linear surrogate importance measure, PC surrogate predictions, total Sobol' indices for PC surrogate. The linear surrogate does not perform as well for this QoI; however, the PC surrogate is far more accurate; this highlights the nonlinearity of this particular QoI. As in the previous results, parameter θ120 appears important in the linear surrogate and less important in the PC surrogate, albeit, it is more important for this QoI than the previous ones. 5 Discussion We note at the outset that the sensitivity analysis investigates the numerical model rather than the physiological one. However, since the numerical model has been validated with experimental data [7, 30], the results indicate parameters (and therefore areas) of importance for physiologists and modellers. The 21 00.10.20.30.40.5Model Value00.10.20.30.40.5Surrogate PredictionLinear SurrogateLinear Surrogate120406080100120140160Parameter Index00.050.10.15Importance Measure00.10.20.30.40.5Model Value00.10.20.30.40.5Surrogate PredictionPC SurrogatePC Surrogate16626375120135136137138139140141142154156157158160Parameter Index00.10.20.30.40.50.6Total Sobol' Index Parameter Identification in Supplementary Material Total Sobol' Index (exp.) Total Sobol' Index (rect.) θ141 θ158 θ139 θ142 z4 in equation (149) ncross in equation (214) z2 in equation (148) z5 in equation (149) θ120 Nominal value 5.5 in equation. (10) 0.6242 0.1239 0.0918 0.0644 0.0571 0.6203 0.1488 0.0954 0.0629 0.0240 Table 3: Five most influential parameters for the [AM + AMp]min QoI when the experimental pulse stimulus is applied. The leftmost column is the parameter, the left-center column identifies the parameter in the Supplementary Material, the right-center column is the total Sobol' index computed for the parameter using the experimental pulse stimulus, and the right column is the total Sobol' index computed for the parameter using the rectangular pulse stimulus. integration times t1, t2 are the start and end of neuronal stimulation. Decreasing or increasing these time does not significantly alter the ranking of the parameters. We discuss below results for each QoI individually. 5.1 Average ECS Potassium It is interesting to note for this particular quantity of interest that the first and third most important parameters are associated not with a K+channel but the persistent Na+channel, NaP. The neuron model used in this analysis was developed from the work of Kager et al [18] and Chang et al [5]. The differential equation governing m4, the activation variable for the NaP channel, is given by equation (4.2). The inactivation of the NaP channel is very small compared to activation and therefore these param- eters make little impact on the extracellular K+. Scatter plots for the average of K+in the ECS against the parameters θ61 and θ62 are shown in Figure 9. For each parameter, there is a linear trend where increasing the parameter yields either an increase or decrease in the K+ECS. The characteristic time for the activation variable for the NaP channel is defined as τ = 1 m4α+m4β which, by using equation (4.2), is constant (6 ms). The scatter plots also indicate this definition. These results show that a variation of 22 Fig. 9: Left: scatter plot of θ62 and the average ECS potassium (2.1); right: scatter plot of θ63 and the average ECS potassium (2.1). ±10% in either θ62 or θ63 can either reduce or increase the extracellular K+by approximately 15%. We should note that these results do not take into account the spatial buffering carried out by the astrocytic syncytium which may have a significantly greater effect on the extracellular K+[4, 21]. The main driver for K+in the extracellular space is the K+Na+ATP-ase pump which has the general form given by (in the dendrite) (cid:16) (N a+ d )3 d + N a+ N a+ d,baseline (cid:17)3 (cid:16) (K + e )2 K + e + K + e,baseline (cid:17)2 . (5.1) The ATP-ase pumps out Na+and in K+in the ratio of 3 to 2, hence the Na+in the dentrite has a large effect on the extracellular K+as seen in equation (5.1) and strengthens the result that the NaP channel has the most effect on the K+. One could have expected the NaT (transient Na+) channel to be prominent however, it has a fast inactivation variable and, although it produces a larger flux, does so over a shorter time. 5.2 Average Volumetric Flow Rate In the presented model, the main pathway for neurovascular coupling is that of the K+pathway. Here the astrocyte takes up K+ from the synaptic cleft and provides an efflux into the perivascular space (PVS) via the BK channel. The smooth muscle cell detects this increase in PVS K+and through the in- wardly rectifying channel, KIR, hypopolarises the SMC, shutting off the voltage mediated Ca2+channel. 23 0.130.140.150.16c155.566.577.58ECS Kmean+55.566.5c255.566.577.58ECS Kmean+ Ca2+is therefore reduced and the SMC dilates. As stated in the Section 4, the most important param- eter is the shift parameter θ141 in the KIR conductance (4.4) determining the magnitude of ion flux per unit change in the membrane potential away from the equilibrium (Nernst potential). Compared to the other parameters defining the KIR conductance, θ141 is large with a nominal value of 12.6 whereas θ140 = 4.2 × 10−4µM−1 and θ142 = −7.4 × 10−2mV −1. Hence, for constant membrane potential and K+, variations in θ141 produce exponentially large variations in the KIR channel conductance allowing substantial efflux of Ca2+from the SMC and a dilation of the vessel(cid:0) dR dt > 0(cid:1). The second most important parameter is the power index for the cytosolic SMC Ca2+which mediates the four state latch model of Hai and Murphy [13] that is used in this model, in particular the rate of phosphorylation of myosin and the actin-myosin complex. Variations in Ca2+as predominantly dictated by the KIR channel will therefore have a direct effect on the dilation/contraction properties of the SMC and hence the perfusing vessel radius. The remaining three parameters have significantly lower total Sobol' indices and therefore only make small contributions to the QoI. 5.3 [AM + AMp]min We see similar parameters appearing as most important for this QoI and the average volumetric flow rate QoI, which is to be expected given the strong relationship between the radius dilation/contraction phenomenon and the total quantity of actin/myosin complex and its phosphorylated compliment. In fact, this QoI serves as a test of the statistical mechanism in that the only two non-repeating parameters (in the list of five most important) between the volumetric flow rate QoI and the actin/myson complex QoI are θ120 and θ136, which, as noted above, are significantly less import than the leading parameters. 6 Conclusion A three stage methodology is presented for global sensitivity analysis of numerical cell models with a large number of parameters. To the authors knowledge, this is the first type of analysis which investigates neuroscience models of such size. The analysis investigated three quantities of interest pertaining to a 24 numerical model of neurovascular coupling. The results indicated several prevalent features of the model. A significant influence of the persistent N a+ channel activation variable on the average extracellular space K +, a two parameter set (the inwardly rectifying K + channel shift parameter and the index for the cytosolic Ca2+) which characterizes most of the variability in the average volumetric flow rate, and strong similarities between the most influential parameters for the average volumetric flow rate and minimum value of the combined actin/myocin complex. In addition to the results reported in this article, four other QoI's were considered. Two of them, the maximum and average potassium concentration in the Astrocyte, were omitted because the parameter to QoI mapping is nearly constant and hence global sensitivity analysis is not necessary. Specifically, the mean of the QoI is approximately 37 times larger than its standard deviation for both cases. The other two unreported QoIs correspond to lag times. The first being the duration of time between the application of the stimulus and the minimum of the phosphorylated actin myosin complex, and the second being the duration of time between the application of the stimulus and the maximum of the radius. In both cases, the QoI exhibited highly nonlinear behavior which we were unable to approximate with linear or PC surrogates trained on the existing data. In fact, fitting such nonlinearities would likely require more samples than is computationally feasible for this model. The linear surrogate had 59% and 47% relative L2 errors for these two QoIs, respectively. Our global sensitivity analysis methodology was unsuccessful because the linear surrogate was an unreliable tool for screening. Defining the QoI as the maximum/minimum value instead of the time lag makes the analysis more tractable. These maximum/minimum value QoIs were also considered and yielded similar results to the QoIs reported in the article. For a given model and collection of samples, the methodology presented in this article may be appli- cable for some QoIs and intractable for others. The success of our method depends upon the surrogate models being sufficiently accurate. A general principle is that QoIs defined as averages will be more amenable for analysis than, for instance, minimum values or lag times. A practical benefit of our method is that any QoI may be considered without requiring additional model evaluations. The sampling and ODE solves are executed once, followed by computing the QoIs and performing global sensitivity analysis, which may be easily repeated for many different QoIs. 25 Appendix Determining the coefficients cα in the Polynomial Chaos surrogate (3.3) is challenging. Ideally, one would solve the least squares problem (cid:32) g(θk) −(cid:88) α M(cid:88) k=1 min (cid:33)2 cαψα( θ) (A.1) to determine the coefficients. This approach is not currently feasible for the problems considered in this article. If there are, for instance, 18 input variables (θji 's) then a 3rd degree polynomial has 1330 unknown coefficients and a 4th degree polynomial has 7315 unknown coefficients. With less than 1000 sample points, as in our case, (A.1) will admit infinitely many solutions which interpolate the data but will yield poor approximations of the QoI. Rather, we seek an approximate solution of (A.1) for which most of the coefficients are exactly 0. This may be achieved by introducing a penalty term and solving M(cid:88) (cid:32) g(θk) −(cid:88) min (cid:33)2 (cid:88) cαψα( θ) + λ cα (A.2) k=1 α α instead of (A.1). Adding the sum of absolute values of the coefficients encourages a sparse solution, i.e. one with many 0 coefficients. However, it comes at the cost of making the objective function non- differentiable and hence (A.2) requires a more sophisticated optimization approach in comparison to (A.1). A plurality of well documented methods exist for solving (A.2). In this article we use Least Angle Regression (LAR) [9] with its implementation in [26], and a maximum polynomial degree of 5. Because the basis function of the Polynomial Chaos surrogate are orthogonal with respect to the PDF p θ, the variance and conditional expectation in (3.6) may be computed analytically as a function of the coefficients. Hence the total Sobol' indices of the Polynomial Chaos surrogate are given in closed form as a function of the coefficients. References 1. Alexanderian A, Gremaud P, Smith R (2018) Variance-based sensitivity analysis for time-dependent processes, under revision, https://arxiv.org/abs/1711.08030 2. Ames A (2000) CNS energy metabolism as related to function. Brain Research Reviews 34:42 -- 68 26 3. Attwell D, Buchan AM, Charpak S, Lauritzen M, MacVicar Ba, Newman Ea (2010) Glial and neuronal control of brain blood flow. Nature 468(7321):232 -- 243 4. Bellot-Saez A, K´ekesi O, Morley JW, Buskila Y (2017) Astrocytic modulation of neuronal excitability through K + spatial buffering. Neuroscience and Biobehavioral Reviews 77:87 -- 97 5. Chang JC, Brennan KC, He D, Huang H, Miura RM, Wilson PL, Wylie JJ (2013) A mathematical model of the metabolic and perfusion effects on cortical spreading depression. PloS One 8(8):e70469 6. Dormanns K, Brown RG, David T (2015) Neurovascular coupling: a parallel implementation. Fron- tiers in Computational Neuroscience 9(September):1 -- 17 7. Dormanns K, van Disseldorp EMJ, Brown RG, David T (2015) Neurovascular coupling and the influence of luminal agonists via the endothelium. Journal of Theoretical Biology 364:49 -- 70 8. Dormanns K, Brown R, David T (2016) The role of nitric oxide in neurovascular coupling. Journal of Theoretical Biology 394:1 -- 17 9. Efron B, Hastie T, Johnstone I, Tibshirani R (2004) Least angle regression. Annals of Statistics 32(11):407 -- 499 10. Farr H, David T (2011) Models of neurovascular coupling via potassium and EET signalling. Journal of Theoretical Biology 286(1):13 -- 23 11. Filosa JA, Bonev AD, Straub SV, Meredith AL, Wilkerson MK, Aldrich RW, Nelson MT (2006) Local potassium signaling couples neuronal activity to vasodilation in the brain. Nature Neuroscience 9(11):1397 -- 1403 12. Filosa JA, Blanco VMVM, Filosa JA Blanco VM, Filosa JA, Blanco VMVM (2007) Neurovascular coupling in the mammalian brain. Experimental Physiology 92(4):641 -- 646 13. Hai CM, Murphy RA (1988) Cross-bridge phosphorylation and regulation of latch state in smooth muscle. American Journal of Physiology - Cell Physiology 254(1):C99 -- C106 14. Hart J, Gremaud P (2018) An approximation theoretic perspective of Sobol' indices with dependent variables, to appear in Int. J. for Uncertainty Quant. 15. Hart J, Alexandrian A, Gremaud P (2017) Efficient computation of Sobol' indices for stochastic models. SIAM J Sci Comput 39:A1514 -- A1530 27 16. Hsieh NH, Reisfeld B, Bois FY, Chiu WA (2018) Applying a global sensitivity analysis workflow to improve the computational efficiencies in physiologically-based pharmacokinetic modeling. Frontiers in Pharmacology 9(588) 17. Iooss B, Saltelli A (2017) Introduction: Sensitivity analysis. In: Ghanem R, Higdon D, Owhadi H (eds) Handbook of Uncertainty Quantification, Springer, pp 1103 -- 1122 18. Kager H, Wadman WJ, Somjen GG (2000) Simulated seizures and spreading depression in a neuron model incorporating interstitial space and ion concentrations. Journal of Neurophysiology 84(1):495 -- 512 19. Kenny A, Plank MJ, David T (2017) The role of astrocytic calcium and TRPV4 channels in neu- rovascular coupling. Journal of Computational Neuroscience doi:10.1007/s10827 -- 017 -- 0671 -- 7, DOI 10.1007/s10827-017-0671-7 20. Kenny A, Plank MJ, David T (2017) The role of astrocytic calcium and TRPV4 channels in neu- rovascular coupling. Journal of Computational Neuroscience (in press) 41(1):97 -- 114 21. Kenny A, Plank M, David T (2018) Macro scale modelling of cortical spreading depression and the role of astrocytic gap junctions. Journal of Theoretical Biology 458:78 -- 91 22. Leithner C, Royl G, Offenhauser N, Fuchtemeier M, Kohl-Bareis M, Villringer A, Dirnagl U, Lindauer U (2010) Pharmacological uncoupling of activation induced increases in CBF and CMRO2. Journal of Cerebral Blood Flow and Metabolism: Official Journal of the International Society of Cerebral Blood Flow and Metabolism 30(2):311 -- 322 23. Lindauer U, Leithner C, Kaasch H, Rohrer B, Foddis M, Fuchtemeier M, Offenhauser N, Steinbrink J, Royl G, Kohl-Bareis M, Dirnagl U (2010) Neurovascular coupling in rat brain operates independent of hemoglobin deoxygenation. Journal of Cerebral Blood Flow & Metabolism 30(4):757 -- 768 24. Longden TA, Hill-Eubanks DC, Nelson MT (2015) Ion channel networks in the control of cerebral blood flow. Journal of Cerebral Blood Flow & Metabolism 36:492 -- 512 25. Makani S, Chesler M (2010) Rapid rise of extracellular pH evoked by neural activity is generated by the plasma membrane calcium ATPase. Journal of Neurophysiology 103(2):667 -- 676 26. Marelli S, Sudret B (2014) UQLab: A framework for uncertainty quantification in Matlab. 2nd Int Conf on Vulnerability, Risk Analysis and Management 28 27. Marin JM, Pudlo P, Robert CP, Ryder RJ (2012) Approximate Bayesian computational methods. Statistics and Computing 22(6):1167 -- 1180 28. Mathias EJ, Plank MJ, David T (2017) A model of neurovascular coupling and the BOLD response: PART I. Computer Methods in Biomechanics and Biomedical Engineering 20(5):508 -- 518 29. Mathias EJ, Plank MJ, David T (2017) A model of neurovascular coupling and the BOLD response: PART II. Computer Methods in Biomechanics and Biomedical Engineering 20(5):519 -- 529 30. Mathias EJ, Kenny A, Plank MJ, David T (2018) NeuroImage Integrated models of neurovascular coupling and BOLD signals: Responses for varying neural activations. NeuroImage 174(February):69 -- 86 31. Mintun MA, Lundstrom BN, Snyder AZ, Vlassenko AG, Shulman GL, Raichle ME (2001) Blood flow and oxygen delivery to human brain during functional activity: Theoretical modeling and ex- perimental data. Proc Natl Acad Sci USA 98(12):6859 -- 6864 32. Owen A (2014) Sobol' indices and Shapley values. SIAM/ASA J Uncertainty Quant 2:245 -- 251 33. Powers WJ, Hirsch IB, Cryer PE (1996) Effect of stepped hypoglycemia on regional cerebral blood flow response to physiological brain activation. Am J Physiol 270(2 Pt 2):H554 -- -9 34. Roy CSCS, Sherrington CSS (1890) On the regulation of the blood-supply of the brain. The Journal of Physiology 11(1-2):85 35. Saltelli A, Ratto M, Andres T, Campolongo F, Cariboni J, Gatelli D, Saisana M, Tarantola S (2008) Global sensitivity analysis: The primer. Wiley 36. Saltelli A, Annoni P, Azzini I, Campolongo F, Ratto M, Tarantola S (2010) Variance based sensitivity analysis of model output. Design and estimator for the total sensitivity index. Computer Physics Communications 181:259 -- 270 37. Sobie EA (2009) Parameter sensitivity analysis in electrophysiological models using multivariable regression. Biophysical Journal 96:1264 -- 1274 38. Tennøe S, Halnes G, Einevoll GT (2018) Uncertainpy: A Python toolbox for uncertainty quantifica- tion and sensitivity analysis in computational neuroscience, http://dx.doi.org/10.1101/274779 39. Witthoft A, Filosa JA, Karniadakis GE (2013) Potassium buffering in the neurovascular unit: models and sensitivity analysis. Biophysical Journal 105(9):2046 -- 2054 29 40. Zheng Y, Pan Y, Harris S, Billings S, Coca D, Berwick J, Jones M, Kennerley A, Johnston D, Martin C, Devonshire IM, Mayhew J (2010) A dynamic model of neurovascular coupling: Implications for blood vessel dilation and constriction. NeuroImage 52(3):1135 -- 1147 30
1907.00263
1
1907
2019-06-29T19:28:25
A Power Efficient Artificial Neuron Using Superconducting Nanowires
[ "q-bio.NC", "cond-mat.supr-con", "cs.NE" ]
With the rising societal demand for more information-processing capacity with lower power consumption, alternative architectures inspired by the parallelism and robustness of the human brain have recently emerged as possible solutions. In particular, spiking neural networks (SNNs) offer a bio-realistic approach, relying on pulses analogous to action potentials as units of information. While software encoded networks provide flexibility and precision, they are often computationally expensive. As a result, hardware SNNs based on the spiking dynamics of a device or circuit represent an increasingly appealing direction. Here, we propose to use superconducting nanowires as a platform for the development of an artificial neuron. Building on an architecture first proposed for Josephson junctions, we rely on the intrinsic nonlinearity of two coupled nanowires to generate spiking behavior, and use electrothermal circuit simulations to demonstrate that the nanowire neuron reproduces multiple characteristics of biological neurons. Furthermore, by harnessing the nonlinearity of the superconducting nanowire's inductance, we develop a design for a variable inductive synapse capable of both excitatory and inhibitory control. We demonstrate that this synapse design supports direct fanout, a feature that has been difficult to achieve in other superconducting architectures, and that the nanowire neuron's nominal energy performance is competitive with that of current technologies.
q-bio.NC
q-bio
A Power Efficient Artificial Neuron Using Superconducting Nanowires Emily Toomey1, Ken Segall2, Karl K. Berggren*1 1Massachusetts Institute of Technology, Department of Electrical Engineering and Computer Science, Cambridge, MA 02139, USA 2Colgate University, Department of Physics, Hamilton, NY 13346, USA *contact: [email protected] ABSTRACT With the rising societal demand for more information-processing capacity with lower power consumption, alternative architectures inspired by the parallelism and robustness of the human brain have recently emerged as possible solutions. In particular, spiking neural networks (SNNs) offer a bio-realistic approach, relying on pulses analogous to action potentials as units of information. While software encoded networks provide flexibility and precision, they are often computationally expensive. As a result, hardware SNNs based on the spiking dynamics of a device or circuit represent an increasingly appealing direction. Here, we propose to use superconducting nanowires as a platform for the development of an artificial neuron. Building on an architecture first proposed for Josephson junctions, we rely on the intrinsic nonlinearity of two coupled nanowires to generate spiking behavior, and use electrothermal circuit simulations to demonstrate that the nanowire neuron reproduces multiple characteristics of biological neurons. Furthermore, by harnessing the nonlinearity of the superconducting nanowire's inductance, we develop a design for a variable inductive synapse capable of both excitatory and inhibitory control. We demonstrate that this synapse design supports direct fanout, a feature that has been difficult to achieve in other superconducting architectures, and that the nanowire neuron's nominal energy performance is competitive with that of current technologies. Keywords: artificial neuron, superconductor, nanowire, spiking neural network, artificial synapse 1. INTRODUCTION The human brain has long been a subject of fascination due to the wide variety of complex operations made possible by groups of a single component -- the neuron. Now, as computation needs are rapidly approaching the limits of traditional von Neumann architectures, the neuron's unique features have led it to become a source of inspiration for new directions for the advancement of computing. Unlike conventional computing schemes, the human brain benefits from characteristics such as extensive parallelism and robustness to errors, allowing it to operate efficiently despite slow speeds on the order of a few Hz. These appealing qualities have spurred the development of technologies that use the brain as a platform for information processing, ranging from small scale modeling of single neuron dynamics to large-scale parallel computing. At the heart of this concept are Spiking Neural Networks (SNNs), which seek to mimic the spiking dynamics of the brain in order to encode information, with the additional benefit of possibly offering new insight into the brain's functionality. In this scheme, spikes serve as the tokens of information, while neurons act analogously to logic gates, producing a single output in response to a combination of multiple inputs [1]. Past approaches to SNNs vary both in degree of bio-realism and in how the spikes are implemented. While software approaches that hard-code spiking dynamics offer flexibility and precision, they are computationally expensive. As a result, other SNNs have used a hardware implementation, relying on devices with intrinsic dynamics that replicate neuron behavior as a means of reducing computation costs. Hardware approaches have been explored in a wide variety of platforms, including CMOS [2], magnetic materials [3], memristors [4] [5], and superconducting Josephson junctions (JJs) [6] [7]. CMOS circuits provide large-scale integration, but are heavy on power dissipation and need many components to achieve biological realism. Memristors and magnetic materials have device characteristics which emulate neural dynamics, but can be slow and are also not energy efficient. Josephson junctions are a fast and energy-efficient technology, but need more components in large networks due to poor fan-out/fan-in properties; furthermore, their weak action potentials are unable to be viewed directly, making diagnostics difficult. These developing technologies highlight several characteristics that are critical for an artificial neuron: (1) inherent device dynamics that are capable of producing spiking behavior; (2) tunable synapses between neurons to allow for the expansion into larger networks with adjustable connectivity; and (3) low power dissipation, both in the dynamic firing state of the neuron and in the static state. As has been emphasized in prior literature[2], optimizing static power dissipation is particularly vital to the goal of creating an energy efficient network. Here we propose an artificial neuron based on superconducting nanowires operating at cryogenic temperatures. Superconductors are prime candidates for generating low-power spiking behavior due to their inherent nonlinearity and negligible static power dissipation. Building on an architecture first implemented in Josephson junctions [6], we rely on the coupling between two shunted nanowires to act analogously to a two-channel neuron for an action potential. We start by describing the nonlinear dynamics of superconducting nanowires, and then present the architecture of the nanowire-based neuron. Using electrothermal circuit simulations, we demonstrate that the device is able to replicate several behaviors of a single neuron. Finally, we present a synapse design and discuss the advantages of the nanowire neuron, including fanout and an energy figure of merit two orders of magnitude better than that of competing technologies, which are critical in moving towards the parallelism of the human brain. 2. THE NANOWIRE NEURON MODEL Superconducting nanowires possess an inherent nonlinearity that serves as the building block of our artificial neuron model. We begin by briefly describing this nonlinearity in a single nanowire, and then present the nanowire neuron circuit and its basic operation principles. 2.1 Relaxation Oscillations The intrinsic nonlinearity of superconducting nanowires makes them ideal candidates for the hardware generation of spiking behavior. When a bias current flowing through a superconducting nanowire exceeds a threshold known as the critical current (Ic), superconductivity breaks down and the nanowire becomes resistive, producing a voltage. The nanowire only switches back to the superconducting state once the bias current is reduced below a level called the retrapping current (Ir), and the resistive portion (the "hotspot") cools down. When the nanowire is placed in parallel with a shunt resistor, this switching process participates in electrothermal feedback with the shunt, producing relaxation oscillations[8]. Relaxation oscillations can be viewed in the context of a simplified action potential. Like the Na+ influx and K+ outflux currents of a neuron, the influx and outflux currents from the nanowire to the shunt resistor are governed by different timescales, τ1 and τ2. As shown in Figure 1a, the rising edge of the output voltage is defined by τ1=L/(Rs+Rhs), where L is the inductance of the nanowire, Rs is the shunt resistance, and Rhs is the resistance of the nanowire hotspot, usually on the order of ~ 1-10 kΩ. Conversely, the outflux current occurs when the nanowire is no longer resistive and the bias current is redirected from the shunt; this reduced resistance results in a slower time constant τ2= L/Rs which defines the falling edge of the output voltage. For typical nanowires devices, τ1 ~ 100 ps and τ2 ~ 1 ns. The two currents are "gated," as shown by the insets in Fig. 1a, by the state of the nanowire -- when the state is resistive (producing a voltage), the influx current flows into the shunt, and when it is superconducting, the outflux current flows back to the nanowire. The inductance of superconducting nanowires is dominated by an intrinsic material property known as kinetic inductance [9], and is defined per unit length of the structure. As a result, it is possible to tune these time constants by changing the length of the nanowire, with a longer wire leading to a higher inductance and thus a longer timescale. Figure 1b shows a scanning electron micrograph of a typical superconducting nanowire with a meandering geometry designed for maximizing the total device inductance. An example of experimentally observed relaxation oscillations for such a device is displayed in Figure 1c. Figure 1: Relaxation oscillations in superconducting nanowires, which serve as the foundation of the nanowire neuron's spiking behavior. (a) Simplified model of a relaxation oscillation. The output voltage is defined by two time constants. The rising edge occurs when the superconducting nanowire has switched into the resistive state (Rhs > 0), and the bias current is redirected to the shunt resistor. The falling edge takes place after the nanowire regains superconductivity (Rhs = 0), and the bias current is redirected from the shunt resistor. (b) Example of a superconducting nanowire in a long, meandered design for obtaining a high kinetic inductance. (c) Experimentally measured relaxation oscillations in a long superconducting nanowire shunted by 50 Ω. 2.2 The Neuron Model Although a shunted nanowire on its own produces oscillations analogous to action potentials, as the bias current increases, the output signal eventually accumulates a voltage offset. This effect deviates from true neuron behavior, as the cell must maintain a constant resting potential (approximately -70 mV). To overcome this difference, we have implemented a neuron architecture based on one that was first proposed for Josephson junctions [6], as shown in Figure 2. The circuit consists of two shunted nanowires -- the main oscillator and the control oscillator -- linked together in a superconducting loop. A bias current Ibias is applied to both oscillators such that they are each biased right below their critical currents, but in opposite directions. To trigger an action potential, a small input current pulse Iin (Fig. 2a) is applied and sums with the bias current to exceed Ic of the main oscillator, causing it to switch (Fig. 2d). The control does not fire since the input opposes the direction of its bias. Once the main oscillator switches, current is added to the superconducting loop in the counterclockwise direction (Fig. 2b), which sums with the bias current to fire the control oscillator (Fig. 2c). The control oscillator removes counterclockwise current from the loop, allowing the main oscillator to fire again. Without the presence of the control oscillator, the main oscillator would only be able to fire once, since the counterclockwise current added to the loop would reduce the total current through the nanowire of the main oscillator below its Ic. The voltage from the main oscillator node (Fig. 2e) serves as the spiking output that is carried down to the next neuron via the synapse. Unlike the output from the single shunted nanowire, the output of the two-nanowire circuit does not accumulate a bias offset, making it a suitable spiking signal. In the context of the two-channel neuron model, the main oscillator acts analogously to the Na+ influx current by adding flux to the superconducting loop in the form of a circulating current. The control oscillator acts analogously to the K+ outflux current by reducing the circulating current, resetting the neuron and allowing the main oscillator to fire again. As described in Ref. [8], the rate at which each oscillator fires depends on the magnitude of the bias current, paralleling the voltage-dependent rate constants of ion gates in the Hodgin-Huxley model [10]. Figure 2: Circuit simulations of the two-nanowire soma, where the two oscillators act analogously to the two ion channels in the simplified neuron model. (a) Input pulse, Iin = 4 µA. (b) Current through the loop inductor. (c) Current through the control nanowire. The control nanowire reduces the amount of counter-clockwise current circulating in the loop, allowing the main nanowire to fire again. (d) Current through the main nanowire. (e) Output voltage pulse that is sent to the synapse. For these simulations, the critical current of the control nanowire is Ic,control = 30 µA, the critical current of the main nanowire is Ic,main = 30 µA, and Ibias = 58.6 µA. 3. SINGLE NEURON CHARACTERISTICS Neurons display a wide variety of traits unique to certain populations, allowing them to collectively interact to achieve varied and complex tasks. While no single neuron possesses all possible traits, the basic functionality of an artificial neuron can be evaluated by demonstrating some common bio-realistic characteristics. Here we present multiple neuron behaviors that can be achieved with the nanowire neuron, using electrothermal circuit simulations conducted in LTSpice. The simulations implement material-specific characteristics and nanowire hotspot dynamics as described in previous literature [11] [12], and have been shown to reliably reproduce experimental data pertaining to nanowire relaxation oscillations [8]. 3.1 Threshold Response A general characteristic of biological neurons is their inability to fire unless the input signal exceeds a certain threshold. Figure 3a shows the threshold voltage response of the nanowire neuron when the bias current is held constant and the input current is varied. As evident in the plot, the neuron does not begin firing until the input current passes a threshold, defined by when the sum of the bias and input current through the main nanowire exceed its Ic. Above the threshold, the peak voltage of the spike output is essentially constant. However, as emphasized by Izhikevich [13], biological neurons have a threshold that may be varied by previous activity, such as an inhibiting input that reduces it. Figure 3b and c illustrate this process ("threshold variability") in the nanowire neuron; an initial subthreshold input pulse (Fig. 3b) fails to elicit a spike, while a later input pulse of the same magnitude triggers a spike (Fig. 3c) after a smaller negative (inhibitory) pulse reduces the firing threshold. It should be noted that when the preceding pulse was of the opposite polarity (excitatory), no spike was triggered. This behavior is consistent with the expectations of Ref. [13]. Figure 3: Firing threshold of the two-nanowire neuron. (a) Peak output voltage as a function of input current under a constant bias (Ibias = 58.6 µA). The plot illustrates that the neuron does not spike until the input current exceeds 4 µA, at which point it fires with output voltages of the same amplitude. Inset shows the time domain voltage output of the neuron for different input currents. (b) Input to the neuron, leading to a reduction in firing threshold by a preceding inhibitory pulse. (c) Spiking output of the neuron in response to the inputs of (b), demonstrating that the nanowire neuron's firing threshold is variable. For this simulation, Ibias = 57.62 µA, the excitatory inputs Iin = 4.6 µA, and the inhibitory input Iin - = -4.3 µA. 3.2 Refractory Period In addition to exhibiting a firing threshold, the nanowire neuron displays a refractory period, which we define to be the minimum time between two input pulses such that both pulses elicit a spike. Figure 4 illustrates this response. When two pulses are separated enough in time so that the main oscillator is biased close to its critical current when the second input pulse arrives, then the second pulse will cause a spike (Fig. 4a). However, if the second pulse arrives before the bias current has fully returned to the main oscillator, then the sum of the second input pulse and the bias will not be sufficient to switch the nanowire and trigger the neuron (Fig. 4b). As a result, the refractory period is limited by the time it takes to fully bias the main oscillator again. Figure 4: Refractory period of the two-nanowire neuron. (a) Response when there is sufficient time between two inputs to each elicit a separate spike. Parameters: Ibias = 58 µA, Iin = 6 µA, Δt = 4 ns. The pink dashed lines indicate the beginning of the rising edge of each pulse. (b) Response when there is insufficient time between two input pulses, causing the neuron to fire only once. Parameters are the same as in (a), except Δt = 2 ns. For both cases, panel (i) displays the current through the nanowire of the main oscillator, while panel (ii) displays the output voltage of the neuron. 3.3 Near-Coincidence Detection The neuron may also be under-biased such that two pulses in rapid succession are required to elicit a spike. Figure 5 displays the response of the nanowire neuron to two input pulses with different time delays Δt between them. In this case, in order for the second pulse to fire the neuron, the delay must be less than 3 ns, as shown in Fig. 5b. This behavior is similar to the response of integrating neurons to high frequency inputs [13], and demonstrates that the nanowire neuron may be used for near-coincidence detection of pulses, depending on its biasing conditions. Figure 5 : Near-coincidence detection of pulses. Output voltage of the neuron when the time between successive input pulses Δt is a) 5 ns b) 3 ns. The pink dashed lines indicate the rising edge of each pulse. Two pulses must be in rapid succession in order to fire the under-biased neuron, demonstrating that it may be used for near-coincidence detection of input pulses. Parameters: Iin = 4.6 µA, Ibias = 57.62 µA. 3.4 Class I Behavior Biological neurons differ in their response to varying signal strengths. Whereas Class I neurons have a spiking frequency that increases with increasing input strength, Class II neurons maintain a constant firing rate[13] [6]. Figure 6 illustrates the spiking behavior of the nanowire neuron at different levels of bias current. Fig. 6a shows the time-domain voltage output of the neuron as the bias current is increased, and suggests an increase in spiking frequency. This response is confirmed by observing the voltage output's frequency spectrum displayed in Fig. 6b, which shows a shift in the spiking frequency to higher levels with increasing bias. Consequently, the nanowire neuron has Class 1 behavior. The modulation of spiking frequency by bias current demonstrates that the frequency of the nanowire neuron output may be used to glean information about its input conditions. Figure 6: Effect of bias current on spiking frequency. (a) Time domain simulations of the two-nanowire neuron with different bias currents. Iin = 6 µA for all simulations. Traces have been shifted from one another in the y-axis for clarity. (b) Fourier transform of the voltage output for each biasing condition. The shift in peak frequency with bias current indicates that the circuit acts like a Class I neuron. 3.5 Parabolic Bursting Some neurons, such as thalamic and dopaminergic neurons [14], display a unique mode of behavior called bursting, in which the cell alternates between the resting and firing states. The transition between states may be dictated by slow changes in low levels of intracellular calcium ions, which influence the conductance of K+ [15]. As a result, a small, slowly varying signal controls the rapid dynamics of the action potential, leading to alternating periods of resting and firing. This process is considered to be an important aspect of electrical activity in the brain. Due to the significance of bursting in biological neurons, past work has sought to replicate similar behavior in platforms such as digital silicon models [16] and Josephson junction models [17] by injecting a low-frequency ac signal into the system. Figure 7 shows the result when a similar technique is applied to the nanowire neuron. In this case, a weak ac signal (f = 50 Hz, Iac = 4 µA) shown by the dashed red line in Fig. 7a is coupled into the bias port of the neuron, causing the neuron to alternate between the resting and firing states, as reflected in the resulting output voltage signal. A close examination of the timing between adjacent spikes (see Fig. 7b) shows that the spiking frequency increases and then decreases over the firing period, a phenomenon known as parabolic bursting[18]. This behavior was first observed in neuron R15 of the abdominal ganglion of Aplysia [19] [20], and has since been demonstrated in many other cells. The ability of the nanowire neuron to replicate similar dynamics may be therefore be useful for performing a wider range of functions. Figure 7: Parabolic bursting in the two-nanowire neuron. (a) Output voltage of the two-nanowire neuron when the bias is coupled to a weak sinusoidal drive (f = 50 MHz, Iac = 4 µA). The red dashed curve indicates the sinusoidal drive, shifted in the y-axis for clarity. (b) The inverse of the time between adjacent peaks shows that the time difference follows a parabolic form. Parameters: Ibias = 59 µA, Iin = 6 µA. 3.6 Axon: Transmission Line Characteristics After an action potential occurs in a biological neuron, the output signal propagates down the axon as if sent through a delay line [1]. This delay is valuable in that it preserves time domain information, potentially facilitating behaviors that rely on the recognition of specific spatio-temporal patterns [13]. Such pattern recognition is often not possible in SNNs with traditional wiring, since signals travel too rapidly for timing information to be maintained [1]. This is not the case in superconducting transmission lines. Recent work has shown that the high kinetic inductance of superconducting transmission lines, like those made out of niobium nitride, results in propagation speeds of ~2% c, where c is the speed of light in vacuum [21]. As illustrated in Figure 8, a simulated nanowire neuron output sent through a superconducting transmission line model [22] is delayed by ~100-500 ps, close to the full width of an action potential. In mammalian brains, the full width of the action potential is typically a few milliseconds [23], and so by analogy the axonal delay shown in Figure 8 would also be on the millisecond timescale. This delay would be on the order of typical delays in cortical systems, for example the cortico- cortical delay in mammals [24]; if longer delays are needed, the transmission line can simply be made longer. Figure 8: A superconducting transmission line as an axon. Simulations of a superconducting transmission line show that the spikes can be delayed on the order of ~0.5 ns, depending on the length of the structure. This could enable the storage of timing information in addition to frequency information. Transmission line parameters: nanowire inductance Ln = 0.3 nH/µm, nanowire capacitance Cn = 0.1 fF/µm, propagation speed v = 1.9%c, transmission line length l = 2.5 mm. Shorter transmission lines on the order of 800 um still had delays of ~140 ps. 4. THE SYNAPSE The collective dynamics of a neural network depend on the ability of a neuron to influence the behavior of another downstream neuron via a synapse. Here we introduce an inductive synapse that can be integrated with the nanowire neuron to facilitate downstream control. We start by demonstrating excitatory and inhibitory control, and then present a scheme for tuning the synaptic strength. 4.1 The Inductive Synapse Figure 9 illustrates the circuit schematic of an inductive synapse that may be implemented in the nanowire neuron. Similar to the slow release of neurotransmitters in response to an action potential, the inductive synapse relies on the slow charging of a large inductor in response to the nanowire neuron's more rapid voltage spikes. The energy stored in the large synapse inductor is then discharged as current into the input port of the target neuron, modulating its behavior. Figure 9: Circuit schematic of the nanowire neuron with an inductive synapse. The output voltage of the neuron charges the large synapse inductor Lsyn. The inductor discharges current into the input port of the downstream target neuron, modulating its behavior. A series resistor Rseries is in place to reduce backaction into the main neuron. The effect of the inductive synapse can be either excitatory or inhibitory based on the sign of the bias current applied to the upstream (main) neuron. These two cases are shown in Figure 10, which displays the voltage outputs of the main and target neurons, as well as the current through the synapse inductor. In the excitatory case (Fig. 10a), the upstream neuron is positively biased such that the inductive synapse discharges a positive current, causing an under-biased target neuron to fire. In the inhibitory case (Fig. 10b), the upstream neuron is negatively biased, causing a negative current to be sent from the synapse to the target, turning off a target that was biased in the firing state. Consequently, the inductive synapse allows for both types of control between neurons. Figure 10: Downstream control of the inductive synapse. (a) Excitatory control. Parameters: Ibias,main = 59 µA, Rseries = 14 Ω, Rsyn,1 = 40 Ω, Lsyn = 0.265 µH, Rsyn,2 = 40 Ω, Ibias,target = 57.17 µA, Iin = 4.6 µA. (b) Inhibitory control. Parameters Ibias,main = -58.6 µA, Rseries = 24 Ω, Rsyn,1 = 45 Ω, Lsyn = 0.23 µH, Rsyn,2 = 40 Ω, Ibias,target = 57.68 µA, Iin = 4.6 µA. For both cases: Panel (i) displays the output voltage of the main neuron, with the red dashed lines indicating the rising and falling edge of the input signal; Panel (ii) displays the current through the synapse inductor; Panel (iii) displays the output voltage of the downstream target neuron. 4.2 Variable Strength Synapse Although the inductive synapse of Figure 10 can be engineered for both excitatory and inhibitory control of a downstream neuron, a fundamental property of artificial neural networks is the ability to modulate that control by adjusting synaptic strength. One possible scheme for implementing such variability in the inductive synapse is to incorporate superconducting nanowires as a different circuit element: a tunable inductor. A nanowire's kinetic inductance increases with increasing bias current, reaching an enhancement of 10-20% near Ic [25]. This modulation has been incorporated into the circuit model of the superconducting nanowire used in these simulations [12]. By placing a high inductance nanowire with an ideal current source in parallel with the synapse inductor, the overall parallel inductance of the synapse can be modulated, which in turn changes the amount of current sent to the target neuron. Figure 11 shows the simulated results for the case of an inhibitory synapse. When a higher modulating current Imod is applied to the nanowire inductor, the overall parallel inductance increases, reducing the amount of current sent to the target. It is important to note that the polarity of the modulating current is not important, since the change in kinetic inductance depends only on the magnitude of the modulating current in relation to its Ic. For instance, Figure 11b illustrates that the modulation in synaptic current for Imod = 5 µA and Imod = -5 µA is roughly the same. As a result, it is clear that the modulation is due to the change in kinetic inductance, and not simply an injection of current by Imod in the opposing direction. Figure 11: Modulating the inductive synapse. (a) Circuit schematic of the inductive synapse with a high- inductance nanowire and ideal current source placed in parallel. L1, L2 << Lnanowire, Lsyn. Rseries, out has been added to further prevent backaction from the target neuron. (b) Simulation of the current through Rseries,out in an inhibitory synapse as a function of different modulation currents. Spikes represent backaction from the firing of the target neuron. (Inset) Enlarged view of the boxed area. Higher modulation currents increase the overall synapse inductance, reducing the amount of current sent to the target. The +/- 5 µA modulation currents have nearly the same effect, showing that modulation is not polarity-dependent. Parameters: Rseries,in = 25 Ω, Rsyn,1 = 39 Ω, Rsyn,2 = 40 Ω, Rseries,out = 0.1 Ω, Lsyn = 0.45 µH, Lnanowire = 0.275 µH, Ic, nanowire = 6 µA, L1 = L2 = 50 pH, , Ibias,target = 57.65 µA, Ibias,main = -59.5 µA In conjunction with the plasticity of synapses, the high degree of parallelism in the brain creates a densely connected, adaptable network able to optimize and adjust for different conditions. An example of fanout in the nanowire neuron is shown in Figure 12. As shown in Fig. 12a, a single neuron is connected to four target neurons through four separate tunable synapses. When each of the four modulating currents is set to Imod = 0, the firing of all four targets is inhibited by the main neuron, as illustrated in Fig. 12b. To weaken the connection with one of the targets (target #4), Imod,4 is set to 8 µA, turning off the inhibiting action on target #4 but allowing it to remain on the three other target neurons, as shown in Fig. 12c. Comparison of the synaptic currents for each of the four targets shows that the synaptic current for target #4 is reduced as a result of the modulated inductance. This modulation illustrates that the nanowire neuron is able to be used in a parallel network where the strength of individual synaptic connections can be adapted. Fan-out may be one area where nanowires will be an improvement over Josephson junctions [6], another superconducting technology with promise for artificial neurons. Both nanowires and Josephson junctions have quantized flux outputs (flux here being defined as the time-integral of the voltage). However, Josephson devices have outputs of only a single flux-quantum, while nanowires have outputs of 50-100 flux quanta (i.e. 70 flux quanta for the device in Figure 2). In pushing the fan-out and fan-in to larger systems, one expects to be eventually limited by parasitic inductances and thermal noises. In such a case, the more substantial signal of the nanowire neuron would permit a larger fan-out and fan-in, leading to a higher degree of parallelism. In addition to the fan-out, the nanowire voltage signals are long enough and large enough to be digitized directly on an oscilloscope, in contrast to their Josephson junction counterparts, allowing for more direct analysis and readout. Finally, there are a suite of three- terminal, power-control devices made from nanowires [26], which could be fabricated alongside the neurons and used to boost signal and match impedances. Figure 12: Fanout of nanowire neuron with a tunable inductive synapse. (a) Simplified circuit model of the fanout circuit. (b) Simulation when Imod = 0 µA for all four target neurons. (i) Output of the main upstream neuron. (ii) Current through each of the four series resistors Rseries out to each of the target neurons. (iii) Output voltage of the four target neurons, shifted on the y-axis for clarity. (c) Simulation when Imod = 0 µA for targets 1-3 and Imod = 8 µA for target #4. (i) Output of the main upstream neuron. (ii) Current through each of the four series resistors Rseries out to each of the target neurons. The current through Rseries,out for target #4 is less than that of the other targets, showing that it has been modulated. (iii) Output voltage of the four target neurons, shifted on the y-axis for clarity. The first three targets have been inhibited, while the inhibitory action on target #4 has been turned off. Parameters: Rseries,in= 7 Ω, Rsyn,1 = Rsyn,2 = 300 Ω, Lsyn = 0.4 µH, Rseries,out = 1 Ω, Lnanowire = 0.275 µH, Ic, nanowire = 6 µA, Ibias,main = -59.5 µA, Ibias,target = 57.65 µA. 4.3 Benchmarking Synaptic Energy and Speed The energy dissipation of both the neuron and the synapse can be calculated using LTSpice by taking the time integral of the current-voltage product of each circuit element. Performing this analysis, we find that the nanowire neuron has an energy of about 0.05 fJ for each action potential, while the synapse is about an order of magnitude less at around 0.005 fJ. In large systems, it will be the synapses that will dominate; typically if there are O(N) neurons there will be O(N2) synapses. Hence, even though the neuron dissipates more energy, it will be the synapses that will dominate the power consumption of a large network. In a spiking neural network, the energy dissipated increases with the speed of the system; spiking twice as often dissipates twice the energy, assuming the energy per spike stays constant. The appropriate figure of merit to compare different technologies is then to take the ratio of speed and power. IBM [27] introduced the figure of merit of synaptic operations per second per watt (SOPS/W). We include a constant factor of about 400 W/W for the nanowire neuron to account for the cryogenic cooling costs. Table 1 compares both the energy per spike and the SOPS/W for the nanowire neuron, the human brain, and two CMOS technologies. We acknowledge that the estimate for the nanowire neuron is a projection from a calculation of a single component, whereas the other entries in the table have actually been measured on large systems. However, given that superconducting platforms have no dissipation in their interconnects, we believe that it is reasonable to project in such a way. In doing so, it is apparent that the nanowire neuron can be a highly competitive technology from a power and speed prospective. Human brain NeuroGrid TrueNorth Nanowire neuron energy/ switch 10 fJ 100 pJ 25 pJ ~0.05 fJ SOPS/Watt 1e14 1e10 4e10 5e14 Table 1: Energy and figure of merit values for different artificial neural networks in comparison to those of the human brain. Energy values for the human brain, NeuroGrid, and TrueNorth are taken from Ref. [28]. 4.4 Discussion on Possible Synapse Alternatives Although the fanout achieved here is far from the level of parallelism in the human brain (where each neuron connects to thousands of neighbors), it suggests that nanowires may serve a unique purpose in the development of future superconducting neural networks, which have thus far struggled to support fanout. Given that nanowires can interface with both CMOS and Josephson junction circuits [29], it may be possible for nanowire neurons to serve as intermediary devices in a network with both platforms. Indeed, a recently proposed neural network with hybrid technologies employed superconducting nanowires as photon detectors, relying instead on optical signals for facilitating high fanout [30] [31]. Although our work uses nanowires solely as electrical components, they can easily be biased to act as photodetectors [32] [33] as well, illustrating that the two different architectures would be compatible for integration. While our work takes more of an electrical approach, the two architectures illustrate the diverse ways in which superconducting nanowires can be used in neural networks, suggesting that a combination of the two technologies may be possible. Furthermore, while the inductive synapse proposed in this work relies on an external modulating bias current, it may be possible to use a superconducting memory cell to store the value of this modulating current, or replace it directly in the circuit. Memory cells based on superconducting nanowire loops have been shown to reliably store states in the form of a trapped circulating current [34] [35] [36], with recent work demonstrating that one can program and store different amounts of current in the loop in discrete quantities [37]. Incorporating these programmable memory cells into the modulating elements of the inductive synapse could offer multi-level storage of synaptic strengths, allowing for more complex applications. It may also be possible for nanowire neurons to connect through alternative synapse designs, such as one based on inductive coupling. Further analysis is needed to verify the possibility of fanout in such a scheme. 5. CONCLUSION By taking advantage of intrinsic nonlinearities in superconducting nanowires, we have developed a platform for a low-power artificial neuron. In this platform, the coupling of two nanowire-based oscillators acts analogously to the two ion channels in a simplified neuron model, producing an output voltage spike that serves as the information-carrying token in these circuits. Using electrothermal circuit simulations, we have shown that the nanowire neuron is able to reproduce universal biological neuron characteristics, such as a firing threshold, as well as unique characteristics distinct to certain neural classes, such as parabolic bursting. Furthermore, we suggested that a nanowire transmission line with a propagation speed of ~ 2% c may be used as an axon delay line, potentially allowing spatio-temporal information to be accessed. These collective behaviors may enable the nanowire neuron to be used in a rich variety of operations. In addition to harnessing the nonlinearity of the nanowire's switching dynamics, we relied on the nonlinearity of the nanowire's kinetic inductance in order to develop a variable inductive synapse. We demonstrated that the total parallel inductance of the synapse can be tuned with a modulating bias current, thereby changing the strength of the signal sent to a downstream target neuron. This scheme proved to be capable of fanout, as demonstrated by a single neuron inhibiting the firing of four target neurons. An energy analysis of this circuit in comparison to other spiking networks illustrated that the nanowire neuron has competitive performance in the dynamic firing state and a figure of merit two orders of magnitude better than certain alternative platforms, while the static state benefits from the lack of power dissipation by the superconducting elements. Although experimental realization of these results is a subject of future work, the analysis performed here suggests that the nanowire neuron is a promising candidate for the advancement of low-power artificial neural networks. Looking forward, networks of superconducting nanowires could be the basis for powerful new computer hardware. Analog blocks of highly connected neurons could be digitally linked to achieve networks with either scale-free or small-worlds connectivity. With superconducting interconnects, entire chips could be wired together with no cost in heat dissipation. The result would be a large-scale neuromorphic processor which could be trained as a spiking neural network to perform tasks like pattern recognition or used to simulate the spiking dynamics of a large, biologically-realistic network. The combination of speed, low power dissipation, and biological realism with only a few components suggests that nanowires could outperform or complement other existing and developing neuromorphic hardware technologies. ACKNOWLEDGEMENTS The authors thank Murat Onen and Brenden Butters for scientific discussions, Akshay Agarwal and Dr. Donnie Keithley for edits, and the entire Quantum Nanostructures and Nanofabrication group. They would also like to acknowledge the generous support of the Bose Foundation. Emily Toomey was supported by the National Science Foundation Graduate Research Fellowship Program (NSF GRFP) under Grant No. 1122374. AUTHOR CONTRIBUTIONS E.T. performed the simulations with input from K.S. and K.K.B. All authors contributed to the design of the circuits and the writing of the paper. SUPPLEMENTARY MATERIAL An example of a capacitive synapse, as well as a demonstration of the tunable inductive synapse in an excitatory case, may be found in the Supplementary Material. DATA AVAILABILITY The nanowire LTSpice model used for this study can be found on GitHub at https://github.com/karlberggren/snspd-spice. All circuit schematic and data files, including the Matlab code used to generate the figures, are available from the authors upon request. Works Cited [1] S. Furber and S. Temple, "Neural systems engineering," J. R. Soc. Interface, vol. 4, no. 13, pp. 193 -- [2] 206, Apr. 2007. I. Sourikopoulos et al., "A 4-fJ/Spike Artificial Neuron in 65 nm CMOS Technology," Front. Neurosci., vol. 11, 2017. [3] S. Li, W. Kang, Y. Huang, X. Zhang, Y. Zhou, and W. Zhao, "Magnetic skyrmion-based artificial neuron device," Nanotechnology, vol. 28, no. 31, p. 31LT01, Jul. 2017. [4] D. Querlioz, O. Bichler, and C. Gamrat, "Simulation of a memristor-based spiking neural network immune to device variations," in The 2011 International Joint Conference on Neural Networks, 2011, pp. 1775 -- 1781. [5] A. Thomas, "Memristor-based neural networks," J. Phys. Appl. Phys., vol. 46, no. 9, p. 093001, Feb. 2013. [6] P. Crotty, D. Schult, and K. Segall, "Josephson junction simulation of neurons," Phys. Rev. E, vol. 82, no. 1, p. 011914, Jul. 2010. [7] M. L. Schneider et al., "Ultralow power artificial synapses using nanotextured magnetic Josephson junctions," Sci. Adv., vol. 4, no. 1, p. e1701329, Jan. 2018. [8] E. Toomey, Q.-Y. Zhao, A. N. McCaughan, and K. K. Berggren, "Frequency Pulling and Mixing of Relaxation Oscillations in Superconducting Nanowires," Phys. Rev. Appl., vol. 9, no. 6, p. 064021, Jun. 2018. [9] A. J. Kerman et al., "Kinetic-inductance-limited reset time of superconducting nanowire photon counters," Appl. Phys. Lett., vol. 88, no. 11, p. 111116, Mar. 2006. [10] G. B. Ermentrout and D. H. Terman, "The Hodgkin -- Huxley Equations," in Mathematical Foundations of Neuroscience, vol. 35, New York, NY: Springer New York, 2010, pp. 1 -- 28. [11] A. J. Kerman, J. K. W. Yang, R. J. Molnar, E. A. Dauler, and K. K. Berggren, "Electrothermal feedback in superconducting nanowire single-photon detectors," Phys. Rev. B, vol. 79, no. 10, p. 100509, Mar. 2009. [12] K. K. Berggren et al., "A Superconducting Nanowire can be Modeled by Using SPICE," Supercond. Sci. Technol., 2018. [13] E. M. Izhikevich, "Which model to use for cortical spiking neurons?," IEEE Trans. Neural Netw., vol. 15, no. 5, pp. 1063 -- 1070, Sep. 2004. [14] R. Bertram, M. J. Butte, T. Kiemel, and A. Sherman, "Topological and phenomenological classification of bursting oscillations," Bull. Math. Biol., vol. 57, no. 3, pp. 413 -- 439, Jan. 1995. [15] J. Rinzel, "Bursting oscillations in an excitable membrane model," in Ordinary and Partial Differential Equations, 1985, pp. 304 -- 316. [16] T. Nanami, K. Aihara, and T. Kohno, "Elliptic and parabolic bursting in a digital silicon neuron model," presented at the 2016 International Symposium on Nonlinear Theory and Its Applications, Yugawara, Japan, 2016, p. 4. [17] S. K. Dana, D. C. Sengupta, and C.- Hu, "Spiking and Bursting in Josephson Junction," IEEE Trans. Circuits Syst. II Express Briefs, vol. 53, no. 10, pp. 1031 -- 1034, Oct. 2006. [18] C. Soto-Treviño, N. Kopell, and D. Watson, "Parabolic bursting revisited," J. Math. Biol., vol. 35, no. 1, pp. 114 -- 128, Nov. 1996. [19] F. Strumwasser, "The demonstration and manipulation of a circadian rhythm in a single neuron," in Circadian clocks, Amsterdam: North-Holland Publishing Co., 1965, pp. 442 -- 462. [20] B. O. Alving, "Spontaneous Activity in Isolated Somata of Aplysia Pacemaker Neurons," J. Gen. Physiol., vol. 51, no. 1, p. 29, Jan. 1968. [21] Q.-Y. Zhao et al., "Single-photon imager based on a superconducting nanowire delay line," Nat. Photonics, vol. 11, no. 4, pp. 247 -- 251, Apr. 2017. [22] Q.-Y. Zhao, D. F. Santavicca, D. Zhu, B. Noble, and K. K. Berggren, "A distributed electrical model for superconducting nanowire single photon detectors," Appl. Phys. Lett., vol. 113, no. 8, p. 082601, Aug. 2018. [23] B. P. Bean, "The action potential in mammalian central neurons," Nat. Rev. Neurosci., vol. 8, p. 451, Jun. 2007. [24] S. Ferraina, M. Paré, and R. H. Wurtz, "Comparison of Cortico-Cortical and Cortico-Collicular Signals for the Generation of Saccadic Eye Movements," J. Neurophysiol., vol. 87, no. 2, pp. 845 -- 858, Feb. 2002. [25] A. J. Annunziata et al., "Tunable superconducting nanoinductors," Nanotechnology, vol. 21, no. 44, p. 445202, Oct. 2010. [26] A. N. McCaughan and K. K. Berggren, "A Superconducting-Nanowire Three-Terminal Electrothermal Device," Nano Lett., vol. 14, no. 10, pp. 5748 -- 5753, Oct. 2014. [27] P. A. Merolla et al., "A million spiking-neuron integrated circuit with a scalable communication network and interface," Science, vol. 345, no. 6197, pp. 668 -- 673, Aug. 2014. [28] S. Furber, "Large-scale neuromorphic computing systems," J. Neural Eng., vol. 13, no. 5, p. 051001, Aug. 2016. [29] Q.-Y. Zhao, A. N. McCaughan, A. E. Dane, K. K. Berggren, and T. Ortlepp, "A nanocryotron comparator can connect single-flux-quantum circuits to conventional electronics," Supercond. Sci. Technol., vol. 30, no. 4, p. 044002, 2017. [30] J. M. Shainline et al., "Circuit designs for superconducting optoelectronic loop neurons," J. Appl. Phys., vol. 124, no. 15, p. 152130, Oct. 2018. [31] J. M. Shainline, "Fluxonic Processing of Photonic Synapse Events," ArXiv190402807 Cs, Apr. 2019. [32] G. N. Gol'tsman et al., "Picosecond superconducting single-photon optical detector," Appl. Phys. Lett., vol. 79, no. 6, pp. 705 -- 707, Aug. 2001. [33] F. Marsili et al., "Single-Photon Detectors Based on Ultranarrow Superconducting Nanowires," Nano Lett., vol. 11, no. 5, pp. 2048 -- 2053, May 2011. [34] A. Murphy, D. V. Averin, and A. Bezryadin, "Nanoscale superconducting memory based on the kinetic inductance of asymmetric nanowire loops," New J. Phys., vol. 19, no. 6, p. 063015, 2017. [35] Q.-Y. Zhao et al., "A compact superconducting nanowire memory element operated by nanowire cryotrons," Supercond. Sci. Technol., vol. 31, no. 3, p. 035009, 2018. [36] A. N. McCaughan, E. A. Toomey, M. Schneider, K. K. Berggren, and S. W. Nam, "A kinetic- inductance-based superconducting memory element with shunting and sub-nanosecond write times," Supercond. Sci. Technol., 2018. [37] E. Toomey, M. Onen, M. Colangelo, B. A. Butters, A. N. McCaughan, and K. K. Berggren, "Bridging the Gap Between Nanowires and Josephson Junctions: A Superconducting Device Based on Controlled Fluxon Transfer," Phys. Rev. Appl., vol. 11, no. 3, p. 034006, Mar. 2019.
1508.00429
1
1508
2015-08-03T14:22:52
A three-threshold learning rule approaches the maximal capacity of recurrent neural networks
[ "q-bio.NC", "cond-mat.dis-nn" ]
Understanding the theoretical foundations of how memories are encoded and retrieved in neural populations is a central challenge in neuroscience. A popular theoretical scenario for modeling memory function is the attractor neural network scenario, whose prototype is the Hopfield model. The model has a poor storage capacity, compared with the capacity achieved with perceptron learning algorithms. Here, by transforming the perceptron learning rule, we present an online learning rule for a recurrent neural network that achieves near-maximal storage capacity without an explicit supervisory error signal, relying only upon locally accessible information. The fully-connected network consists of excitatory binary neurons with plastic recurrent connections and non-plastic inhibitory feedback stabilizing the network dynamics; the memory patterns are presented online as strong afferent currents, producing a bimodal distribution for the neuron synaptic inputs. Synapses corresponding to active inputs are modified as a function of the value of the local fields with respect to three thresholds. Above the highest threshold, and below the lowest threshold, no plasticity occurs. In between these two thresholds, potentiation/depression occurs when the local field is above/below an intermediate threshold. We simulated and analyzed a network of binary neurons implementing this rule and measured its storage capacity for different sizes of the basins of attraction. The storage capacity obtained through numerical simulations is shown to be close to the value predicted by analytical calculations. We also measured the dependence of capacity on the strength of external inputs. Finally, we quantified the statistics of the resulting synaptic connectivity matrix, and found that both the fraction of zero weight synapses and the degree of symmetry of the weight matrix increase with the number of stored patterns.
q-bio.NC
q-bio
A three-threshold learning rule approaches the maximal capacity of recurrent neural networks Alireza Alemi1,2,*, Carlo Baldassi1,2, Nicolas Brunel3, Riccardo Zecchina1,2 1 Human Genetics Foundation (HuGeF), Turin, Italy 2 DISAT, Politecnico di Torino, Turin, Italy 3 Departments of Statistics and Neurobiology, University of Chicago, USA * [email protected] Abstract Understanding the theoretical foundations of how memories are encoded and retrieved in neural populations is a central challenge in neuroscience. A popular theoretical scenario for modeling memory function is the attractor neural network scenario, whose prototype is the Hopfield model. The model simplicity and the locality of the synaptic update rules come at the cost of a poor storage capacity, compared with the capacity achieved with perceptron learning algorithms. Here, by transforming the perceptron learning rule, we present an online learning rule for a recurrent neural network that achieves near-maximal storage capacity without an explicit supervisory error signal, relying only upon locally accessible information. The fully-connected network consists of excitatory binary neurons with plastic recurrent connections and non-plastic inhibitory feedback stabilizing the network dynamics; the memory patterns to be memorized are presented online as strong afferent currents, producing a bimodal distribution for the neuron synaptic inputs. Synapses corresponding to active inputs are modified as a function of the value of the local fields with respect to three thresholds. Above the highest threshold, and below the lowest threshold, no plasticity occurs. In between these two thresholds, potentiation/depression occurs when the local field is above/below an intermediate threshold. We simulated and analyzed a network of binary neurons implementing this rule and measured its storage capacity for different sizes of the basins of attraction. The storage capacity obtained through numerical simulations is shown to be close to the value predicted by analytical calculations. We also measured the dependence of capacity on the strength of external inputs. Finally, we quantified the statistics of the resulting synaptic connectivity matrix, and found that both the fraction of zero weight synapses and the degree of symmetry of the weight matrix increase with the number of stored patterns. Author Summary Recurrent neural networks have been shown to be able to store memory patterns as fixed point attractors of the dynamics of the network. The prototypical learning rule for storing memories in attractor neural networks is Hebbian learning, which can store up to 0.138N uncorrelated patterns in a recurrent network of N neurons. This is very far from the maximal capacity 2N , which can be achieved by supervised rules, e.g. by the perceptron learning rule. However, these rules are problematic for neurons in the neocortex or the hippocampus, since they rely on the 1 computation of a supervisory error signal for each neuron of the network. We show here that the total synaptic input received by a neuron during the presentation of a sufficiently strong stimulus contains implicit information about the error, which can be extracted by setting three thresholds on the total input, defining depression and potentiation regions. The resulting learning rule implements basic biological constraints, and our simulations show that a network implementing it gets very close to the maximal capacity, both in the dense and sparse regimes, across all values of storage robustness. The rule predicts that when the total synaptic inputs goes beyond a threshold, no potentiation should occur. Introduction One of the fundamental challenges in neuroscience is to understand how we store and retrieve memories for a long period of time. Such long-term memory is fundamental for a variety of our cognitive functions. A popular theoretical framework for storing and retrieving memories in recurrent neural networks is the attractor network model framework [1–3]. Attractors, i.e. stable states of the dynamics of a recurrent network, are set by modification of synaptic efficacies in a recurrent network. Synaptic plasticity rules specify how the efficacy of a synapse is affected by pre- and post-synaptic neural activity. In particular, Hebbian synaptic plasticity rules lead to long-term potentiation (LTP) for correlated pre- and post-synaptic activities, and long-term depression (LTD) for anticorrelated activities. These learning rules build excitatory feedback loops in the synaptic connectivity, resulting in the emergence of attractors that are correlated with the patterns of activity that were imposed on the network through external inputs. Once a set of patterns become attractors of a network (in other words when the network “learns” the patterns), upon a brief initial activation of a subpopulation of neurons, the network state evolves towards the learned stable state (the network “retrieves” a past stored memory), and remains in that state after removal of the external inputs (and hence maintains the information in short-term memory). The set of initial network states leading to a memorized state is called the basin of attraction, whose size determines how robust a memory is. The attractor neural network scenario was originally explored in networks of binary neurons [1, 2], and then extended from the 90s to networks of spiking neurons [4–7]. Experimental evidence in different areas of the brain, including inferotemporal cortex [8–11] and prefrontal cortex [12–14], has provided support for the attractor neural network framework, using electrophysiological recordings in awake monkeys performing delayed response tasks. In such experiments, the monkey has to maintain information in short-term (working) memory in a ‘delay period’ to be able to perform the task. Consistent with the attractor network scenario, some neurons exhibit selective persistent activity during the delay period. This persistent activity of ensembles of cortical neurons has thus been hypothesized to form the basis of the working memory of stimuli shown in these tasks. One of the most studied properties of attractor neural network as a model of memory is its storage capacity, i.e. how many random patterns can be learned in a recurrent network of N neurons in the large N limit. Storage capacity depends both on the network architecture and on the synaptic learning rule. In particular, the Hopfield network [1] that uses a Hebbian learning rule has a storage capacity of 0.138N in the limit of N → ∞ [15]. Later studies showed how the capacity depends on the connection probability in a randomly connected network [16, 17] and on the coding level (fraction of active neurons in a pattern) [18, 19]. A natural question is, what is the maximal capacity of a given network architecture, over all possible learning rules? This question was answered by Elizabeth Gardner, who showed that the capacity of fully connected networks of In many models, the storage capacity scales with N . 2 binary neurons with dense patterns scales as 2N [20], a storage capacity which is much larger than the one of the Hopfield model. The next question is what learning rules are able to saturate the Gardner bound? A simple learning rule that is guaranteed to achieve this bound is the perceptron learning rule (PLR) [21] applied to each neuron independently. However, unlike the rule used in the Hopfield model, the perceptron learning rule is a supervised rule that needs an explicit “error signal” in order to achieve the Gardner bound. While such an error signal might be available in the cerebellum [22–24], it is unclear how error signals targeting individual neurons might be implemented in cortical excitatory synapses. Therefore, it remains unclear whether and how networks with realistic learning rules might approach the Gardner bound. The goal of the present paper is to propose a learning rule whose capacity approaches the maximal capacity of recurrent neural networks by transforming the original perceptron learning rule such that the new rule does not explicitly use an error signal. The perceptron learning rule modifies the synaptic weights by comparing the desired output with the actual output to obtain an error signal, subsequently changing the weights in the opposite direction of the error signal. We argue that the total synaptic inputs (‘local fields’) received by a neuron during the presentation of a stimulus contain some information about the current error (i.e. whether the neuron will end up in the right state after the stimulus is removed). We use this insight to build a field dependent learning rule that contains three thresholds separating no plasticity, LTP and LTD regions. This rule implements basic biological constraints: (a) it uses only information local to the synapse; (b) the new patterns can be learned incrementally, i.e. it is an online rule; (c) it does not need an explicit error signal; (d) synapses obey Dale’s principle, i.e. excitatory synapses are not allowed to have negative weights. We studied the capacity and the size of the basins of attraction for a binary recurrent neural network in which excitatory synapses are endowed with this rule, while a global inhibition term controls the global activity level. We investigated how the strength of external fields and the presence of correlations in the inputs affect the memory capacity. Finally, we investigated the statistical properties of the connectivity matrix (distribution of synaptic weights, degree of symmetry). Results The network We simulated a network of N binary (McCulloch-Pitts) neurons, fully-connected with excitatory synapses (Fig 1A). All the neurons feed a population of inhibitory neurons which is modeled as a single aggregated inhibitory unit. This state-dependent global inhibition projects back onto all the neurons, stabilizing the network and controlling its activity level. At each time step, the activity (or the state) of neuron i (i = 1 . . . N ) is described by a binary variable si ∈ {0, 1}. The state is a step function of the local field vi of the neuron: si = Θ (vi − θ) , (1) where Θ is the Heaviside function (Θ (x) = 1 if x > 0 and 0 otherwise) and θ is a neuronal threshold. The local field vi represents the overall input received by the neuron from its excitatory and inhibitory connections (Fig 1B). The excitatory connections are of two kinds: recurrent connections from within the excitatory population, and external inputs. The recurrent excitatory connections are mediated by synaptic weights, denoted by a matrix W whose elements wij (the weight of the synapse from neuron j to i) are continuous non-negative variables (wij ∈ [0,∞); wii = 0). In the following, and in all our simulations, we assume that the weights are initialized randomly before the training takes place (see Materials and Methods). 3 Figure 1: A sketch of the network and the neuron model. A. Structure of the network. The fully-connected network consists of N binary (si ∈ {0, 1}) neurons and an aggregated in- hibitory unit. The global inhibition is a function of the state of the network and the external fields, i.e. I(~x, ~s). A memory pattern ~ξ is encoded as strong external fields, i.e. ~x = X ~ξ and presented to the network during the learning phase. B. Each neuron receives excitatory recurrent inputs (thin black arrows) from the other neurons, a global inhibitory input (red connections), and a strong binary external field (xi ∈ {0, X}; thick black arrows). All these inputs are summed to obtain the total field, which is then compared to a neuronal threshold θ; the output of the neuron is a step function of the result. 4 Therefore, in the absence of external inputs, the local field of each neuron i is given by: N vi = Xj=1 where I0 (~s) represents the inhibitory input. For the sake of simplicity, we simulated a synchronous update process, in which the activity of each neuron si is computed from the local field vi at the previous time step, and all updates happen in parallel. wijsj − I0 (~s) , (2) The network was designed so that, in absence of external input and prior to the training process, it should spontaneously stabilize itself to some fixed overall average activity level f (fraction of active neurons, or sparseness), regardless of the initial conditions. In particular, we aimed at avoiding trivial attractors (the all-off and all-on states). To this end, we model the inhibitory feedback (in absence of external inputs) as a linear function of the overall excitatory activity: I0(~s) = H0 + λ N Xi=1 si − f N! . (3) the excitatory network has the desired activity level f , i.e. when PN The parameters H0 and λ can be understood as follows: H0 is the average inhibitory activity when i=1 si = f N ; λ measures the strength of the inhibitory feedback onto the excitatory network. This expression can be interpreted as a first-order approximation of the inhibitory activity as a function of the excitatory activity around some reference value f N , which is reasonable under the assumption that the deviations from f N are small enough. Indeed, by properly setting these two parameters in relation to the other network parameters (such as θ and the average connection strength) it is possible to achieve the desired goal of a self-stabilizing network. i ∈ {0, 1}), and assume that each entry ξµ In the training process, the network is presented a set of p patterns in the form of strong external inputs, representing the memories which need to be stored. We denote the patterns as {~ξµ} (where µ = 1...p and ξµ is drawn randomly and independently. For simplicity, the coding level f for the patterns was set equal to the ξµ i = 1 with probability f , 0 otherwise. During spontaneous activity level of the network, i.e. the presentation of a pattern µ, each neuron i receives an external binary input xi = Xξµ i , where X denotes the strength of the external inputs, which we parameterized as X = γ√N . In addition, the external input also affects the inhibitory part of the network, eliciting a response which indirectly downregulates the excitatory neurons. We model this effect as an additional term H1 in the expression for the inhibitory term (Eq. 3), which therefore becomes: i I(~x, ~s) = H0 + H1PN i=1 xi f N X N + λ( Xi=1 si − f N ), The general expression for the local field vi then reads: vi = N Xj=1 wij sj + xi − I(~x, ~s). (4) (5) In the absence of external fields, xi = 0 for all i, and thus Eqs. 4 and 5 reduce to the previous expressions Eqs. 3 and 2. The goal of the learning process is to find values of wij ’s such that the patterns {~ξµ} become attractors of the network dynamics. Qualitatively, this means that, if the training process is 5 i=1 ξµ i.e. whenever the Hamming distance d = PN successful, then whenever the network state gets sufficiently close to one of the stored patterns, i − si between the current network state and a pattern µ is sufficiently small, the network dynamics in the absence of external inputs should drive the network state towards a fixed point equal to the pattern itself (or very close to it). The general underlying idea is that, after a pattern is successfully learned, some brief external input which initializes the network close to the learned state would be sufficient for the network to recognize and retrieve the pattern. The maximum value of d for which this property holds is then called the basin of attraction size (or just basin size hereafter for simplicity); indeed, there is generally a trade-off between the number of patterns which can be stored according to this criterion and the size of their basin of attraction. More precisely, the requirement that a pattern ~ξµ is a fixed point of the network dynamics in the absence of external fields can be reduced to a condition for each neuron i (cfr. Eqs. 4 and 5): N ∀i : Θ Xj=1  wij ξµ j − I(cid:16)~0, ~ξµ(cid:17) − θ i .  = ξµ (6) This condition only guarantees that, if the network is initialized into a state ~s = ~ξµ, then it will not spontaneously change its state, i.e. it implements a zero-size basin of attraction. A simple way to enlarge the basin size is to make the requirement in Eq. 6 more stringent, by enforcing a more stringent constraint for local fields: ∀i :   PN PN j=1 wij ξµ j=1 wij ξµ j − I(cid:16)~0, ~ξµ(cid:17) > θ + f√N ǫ j − I(cid:16)~0, ~ξµ(cid:17) < θ − f√N ǫ if ξµ if ξµ i = 1 i = 0, (7) where ǫ ≥ 0 is a robustness parameter. When ǫ = 0, we recover the previous zero-basin-size scenario; increasing ǫ we make the neurons’ response more robust towards noise in their inputs, and thus we enlarge the basin of attraction of the stored patterns (but then fewer patterns can be stored, as noted above). The three-threshold learning rule (3TLR) In the training phase, the network is presented with patterns as strong external fields xi. Patterns are presented sequentially in random order. For each pattern µ, we simulated the following scheme: Step 1: The pattern is presented (i.e. the external inputs xi are set to Xξµ i ). A single step of synchronous updating is performed (Eqs. 1, 4 and 5). If the external inputs are strong enough, i.e. γ is large enough, this updating sets the network in a state corresponding to the presented pattern. Step 2: Learning occurs. Each neuron i may update its synaptic weights depending on 1) their current value wt ij , 2) the state of the pre-synaptic neurons, and 3) the value of the local field vi. Therefore, all the information required is locally accessible, and no explicit error signals are used. The new synaptic weights wt+1 are set to: ij wt ij − ηsj, wt ij + ηsj, wt ij , if θ0 < vi < θ if θ < vi < θ1 otherwise, wt+1 ij =  6 (8) where η is the learning rate, and θ0 and θ1 are two auxiliary learning thresholds set as θ0 = θ − (γ + ǫ) f√N θ1 = θ + (γ + ǫ) f√N . (9) (10) We refer to this update scheme as “three-threshold learning rule” (3TRL). After some number of presentations, we checked whether the patterns are learned by presenting a noisy version of these patterns, and checking whether the patterns (or network states which are very close to the patterns) are fixed points of the network dynamics. When N ≫ 1, γ is large enough, and H1 = f X, the update rule described by Eq. 8 is essentially equivalent to the perceptron learning rule for the task described in Eq. 7. This can be shown as follows (see also Fig 2 for a graphical representation of the case f = 0.5 and ǫ = 0): when a stimulus is presented, the population of neurons is divided in two groups, one for which xi = 0 and one for which xi = X. The net effect of the stimulus presentation on the local field has to take into account the indirect effect through the inhibitory part of the network (see Eq. 4), and thus is equal to −f X for the xi = 0 population and to (1 − f ) X for the xi = X population. Before learning, the distribution of the local fields across the excitatory population, in the limit N → ∞, is a Gaussian whose standard deviation is proportional to √N , due to the central limit theorem; moreover, the parameter H0 is set so that the average activity level of the network is f , which means that the center of the Gaussian will be within a distance of order √N from the neuronal threshold θ (this also applies if we use different values for the spontaneous activity level and the pattern activity level). Therefore, if X = γ√N is large enough, the state of the network during stimulus presentation will be effectively clamped to the desired output, i.e. si = ξµ for all i. This fact has two consequences: 1) the local field potential can be used to i detect the desired output by just comparing it to the threshold, and 2) each neuron i will receive, . Furthermore, due to the choice of the secondary thresholds θ0 and θ1 in Eqs. 9 and 10, the difference between the local field and θ0 (or θ1) during stimulus presentation for the xi = 0 population (or xi = X, respectively) as its recurrent inputs {sj}j6=i, the rest of the pattern (cid:8)ξµ is equal to the difference between the local field and θ − f√N ǫ (or θ + f√N ǫ, respectively) in j(cid:9)j6=i the absence of external stimuli, provided the recurrent inputs are the same. Therefore, the value of the local field vi during stimulus presentation in relation to the three thresholds θ, θ0 and θ1 is sufficient to determine whether an error is made with respect to the constraints of Eq. 7, and which kind of error is made. Following these observations, it is straightforward to map the standard perceptron learning rule on the 4 different cases which may occur (see Fig 2), resulting in Eq. 8. In Fig 3 we demonstrate the effect of the learning rule on the distribution of the local field potentials as measured from a simulation (with f = 0.5 and ǫ = 1.2): the initial distribution of the local fields of the neurons, before the learning process takes place and in the absence of external fields, is well described by a Gaussian distribution centered on the neuronal threshold θ (see Fig 3A) with a standard deviation which scales as √N . During a pattern presentation, the resulting distribution becomes a bimodal one; before learning takes place, the distribution is given by the sum of two Gaussians of equal width, centered around θ0 + f√N ǫ and θ1 − f√N ǫ (Fig 3B). The left Gaussian corresponds to the cases where xi = 0 and the right one to the cases where xi = X. Having applied the learning rule, we observe that the depression region (i.e. the interval (θ0, θ)) and the potentiation region (i.e. (θ, θ1)) gets depleted (Fig 3C). In the testing phase, when the external inputs are absent, the left and right parts of the distribution come closer, such that the distance between the two peaks is equal to at least 2ǫf√N (Fig 3D). This margin between the local fields of the ON and OFF neurons makes the attractors more robust. 7 Original perceptron rule 3TLR d e r i s e d ) d ( t u p t u o l a u t c a ) y ( t u p t u o 0 0 1 1 0 1 0 1 l i a n g s r o r r e ) y - d ( 0 -1 +1 0 Δw Ø D P Ø d = 0 y = 0 do nothing y = 1 depress active w v d = 1 y = 0 potentiate active w y = 1 do nothing v xi = 0 xi = X do nothing depress active w potentiate active w do nothing v Figure 2: The three-threshold learning rule (3TLR), and its relationship with the stan- dard perceptron learning rule (PLR). The perceptron learning rule modifies the synaptic weights by comparing the desired output with the actual output to obtain an error signal, subse- quently changing the weights in the opposite direction of the error signal (see the table in the left panel). For a pattern which is uncorrelated with the current synaptic weights, the distribution is Gaussian (in the limit of large N ), due to the central limit theorem. H0 is set such that, on average, a fraction f of the local fields are above the neuronal threshold θ; in the case of f = 0.5, this means that the Gaussian is centered on θ (left panel). In our model (Fig 1B), the desired output is given as a strong external input, whose distribution across the population is bimodal (with two delta functions on xi = 0 and xi = X); therefore, the distribution of the local fields during stimulus presentation becomes bimodal as well (right panel). The left and right bumps of this distribution correspond to cases where the desired outputs are zero and one, respectively. Note that, since the external input also elicits an inhibitory response, the neurons in the net- work which are not directly affected by the external input (i.e. those with desired output equal to zero) are effectively hyperpolarized. If X is sufficiently large, the two distributions do not overlap, and the four cases of the PLR can be mapped to the four regions determined from the three thresholds, indicated by vertical dashed lines (see text). 8 WITHOUT EXTERNAL INPUT WITH EXTERNAL INPUT ON neurons OFF neurons B C Total input A y t i s n e d y t i l i b a b o r P D I G N N R A E L E R O F E B I G N N R A E L R E T F A Figure 3: Distribution of local fields before and after learning for f = 0.5 and non-zero robustness. A. Before learning begins, the distribution of local field of neurons is a Gaussian distribution (due to central limit theorem) centered around neuronal threshold θ both for neurons with the desired output zero (OFF neurons) and with the desired output one (ON neurons). The goal is to have the local field distribution of ON neurons (red curve) to be above the threshold θ, and that of OFF neurons to be below θ. B. Once any of the to-be-stored patterns are presented as strong external fields, right before the learning process starts, the local field distribution of the OFF neuron shifts toward the left-side centered around θ0+f ǫ√N , whereas the distribution of the ON neurons moves toward the right-side, centered around θ1 − f ǫ√N , with a negligible overlap between the two curves if the external field is strong enough. Thanks to the strong external fields and global inhibition, the local fields of the ON and OFF neurons are well separated. C. Due to the learning process, the local fields within the depression region [i.e. (θ0, θ)] get pushed to the left-side, below θ0, whereas those within the potentiation region get pushed further to the right-side, above θ1. If the learning process is successful, it will result in a region (θ0, θ1) which no longer contain local fields, with two sharp peaks on θ0 and θ1. D. After successful learning, once the external fields are removed, the blue and red curves come closer, with a gap equal to 2f ǫ√N . The larger the robustness parameter ǫ, the more the gap between the left- and right-side of the distribution. Notice that now the red curve is fully above θ which means those neurons remain stably ON, while the the blue curve is fully below θ, which means those neurons are stably OFF. Therefore the corresponding pattern is successfully stored by the network. Storage capacity Since our proposed learning rule is able to mimic (or approximate, depending on the parameters) the perceptron learning rule, which is known to be able to solve the task posed by Eq. 7 whenever 9 a solution exists, we expect that a network implementing such rule can get close to maximal capacity in terms of the number of memories which it can store at a given robustness level. The storage capacity, denoted by α = p/N , is measured as a ratio of the maximum number of patterns p which can successfully be stored to the number of neurons N , in the limit of large N . As mentioned above, it is a function of the basin size. We used the following definition for the basin size: a set of p patterns is said to be successfully stored at a size b if, for each pattern, the retrieval rate when starting from a state in which a fraction b of the pattern was randomized is at least 90%. The retrieval rate is measured by the probability that the network dynamics is able to bring the network state to an attractor within 1% distance from the pattern, in at most 30 steps. The distance between the state of the network and a pattern µ is measured by the normalized Hamming distance 1 i . Therefore, at coding level f = 0.5, reaching a basin size b means that the network can successfully recover patterns starting from a state at distance b/2. i=1 si − ξµ N PN Fig 4A shows the maximal capacity as a function of the basin size for a simulated network of N = 1001 neurons. We simulated many pairs of (α, ǫ) with different random seeds, obtaining a probability of success for each pair. The red line shows the points for which the probability of successful storage is 0.5, and the error bars span 0.95 to 0.05 success probability. The capacity was optimized over the robustness parameter ǫ. The maximal capacity (the Gardner bound) in the limit of N → ∞ at the zero basin size is αc = 2 for our model (see Materials and Methods for the calculation), as for a network with unconstrained synaptic weights [20]. In Fig 4A, we also compare our network with the Hopfield model. Our network stores close to the maximal capacity at zero basin size, at least eleven times more than the Hopfield model. Across the range of basin sizes, 3TLR achieves more than twice the capacity that can be achieved with the Hopfield model. The enlargement of the basin of attraction was achieved by increasing the robustness pa- rameter ǫ. We computed the maximal theoretical capacity as a function of ǫ at N → ∞ (see Materials and Methods) and compared it to our simulations, and to the maximal theoretical capacity of the Hopfield network. The results are shown in Fig 4B. For any given value of ǫ, the cyan curve shows the maximum α for which the success ratio with our network was at least 0.5 across different runs. The difference between the theory and the experiments in our model can be ascribed to several factors: the finite size of the network; the choice of the finite learning rate η, and the fact that we imposed a hard limit on the number of pattern presentations (see number of iterations in Table 1), while the perceptron rule for excitatory synaptic connectivity is only guaranteed to be optimal in the limit of η → 0, with a number of presentations inversely proportional to η [25]. Note that the correspondence between the PLR and the 3TLR is only perfect in the large γ limit, and is only approximate otherwise, as can be shown by comparing explicitly the synaptic matrices obtained by both algorithms on the same set of patterns (see Materials and Methods.) A crucial ingredient of the 3TLR is having a strong external input which effectively acts as a supervisory signal. How strong do the external fields need to be? How much does the capacity depend on this strength? To answer these questions, we measured the maximum number of stored patterns as a function of the parameter γ which determines the strength of external fields as X = γ√N . This parameter, in fact, determines how far the two Gaussian distributions of the local field are; as shown in Fig 2, the distance between the two peaks of the distribution is X. For large enough γ, the overlap of these two distributions is negligible and the capacity is maximal; but as we lower γ, the overlap increases, causing the learning rule to make mistakes, i.e. when it should potentiate, it depresses the synapses and vice versa. In our simulations with N = 1001 neurons in the dense regime f = 0.5 at a fixed epsilon ǫ = 0.3, we varied γ and computed the maximum α that can be achieved with a fixed number of iterations (1000). The capacity indeed gradually decreases as γ decreases, until it reaches a threshold, below which there is a sharp 10 A B 2 Gardner bound Our model Hopfield model ) c (cid:4) ( y t i c a p a C 1.5 1 0.5 0 0 2 ) c (cid:4) ( y t i c a p a C 1.5 1 0.5 0 0 f = 0.5 (cid:1) = 6 0.2 0.4 0.6 Basin size 0.8 1 Gardner (bounded synapses) Gardner (unbounded synapses) Simulation (3TLR) f = 0.5 (cid:1) = 6 1 Robustness((cid:3)) 2 3 Figure 4: Critical capacity as a function of the basin size and the robustness param- eter. A. The red plot shows the critical capacity as a function of the size of the basins of attraction (N = 1001 neurons in the dense regime f = 0.5) when the strength of the external field is large (γ = 6) such that the ON and OFF neuronal populations are well separated. The points indicate 0.5 probability of successful storage at a given basin size, optimized over the robustness parameter ǫ . The error bars show the [0.95, 0.05] probability interval for successful storage. The blue plot shows the performance of the Hopfield model with N = 1001 neurons. The maximal capacity at zero basin size (the Gardner bound) is equal to 2. B. To compare the result of simulation of our model with the analytical results, we plotted the critical capacity as a function of the robustness parameter ǫ. The dark red curve is the critical capacity versus ǫ for our model obtained form analytical calculations (see Materials and Methods), the cyan line shows the result of simulations of our model, and the dark blue shows the Gardner bound for a network with no constraints on synaptic weights. The difference between the two theoretical curves is due to the constraints on the weights in our network. drop of capacity (see Fig 5). With the above values for the parameters, this transition occurs at γ ≈ 2.4. 11 1 0.5 (cid:1) m u m i x a M 0 0 f = 0.5 (cid:4) = 0.3 2 (cid:3) 4 6 Figure 5: Dependence of the critical capacity on the strength of the external input. We varied the strength of the external field (γ) in order to quantify its effect on the learning process. The critical capacity is plotted as a function of γ at a fixed robustness ǫ = 0.3 in the dense regime f = 0.5. The simulations show that there is a very sharp drop in the maximum α when γ goes below ≈ 2.4. 3 2.5 2 1.5 1 0.5 ) c (cid:1) ( y t i c a p a C 0 0 Theory Simulation f = 0.2 (cid:4) = 12 1 Robustness ((cid:2)) 2 3 Figure 6: Capacity as a function of the robustness parameter ǫ at sparseness f = 0.2. The theoretical calculations is compared with the simulations for f = 0.2. Note that the capacity in the sparse regime is higher than in the dense regime. The 3TLR can also be adapted to work in a sparser regime, at a coding level lower than 0.5. However, the average activity level of the network is determined by H0, and their relationship also involves the variance of the distribution of the synaptic weights when f 6= 0.5 (see Materials and Methods). During the learning process, the variance of the weights changes, which implies that the parameter H0 must adapt correspondingly. In our simulations, this adaptation was performed after each complete presentation of the whole pattern set. In practice, this additional 12 self-stabilizing mechanism could still be performed in an unsupervised fashion along with (or in alternation with) the learning process. Using this adjustment, we simulated the network at f = 0.2 and compared the results with the theoretical calculations. As shown in Fig 6, we can achieve at least 70% of the critical capacity across different values of the robustness parameter ǫ. f = 0.2 (cid:2) = 12 (cid:3) = 3 ) c (cid:4) ( 1 0.8 0.6 0.4 0.2 0 0 0.25 0.5 0.75 ) Figure 7: Capacity as a function of correlations in the input patterns, for f = 0.2 at ǫ = 3.0. Patterns are organized in categories, with a correlation c with the prototype of the corresponding category (see text). We also investigated numerically the effect of correlations in the input patterns. The PLR is able to learn correlated patterns as long as a solution to the learning problem exists. As the 3TLR approximates the PLR, we expect the 3TLR to be able to learn correlated patterns as well. As a simple model of correlation, we tested patterns organized in L categories [26,27]. Each category was defined by a randomly generated prototype. Prototypes were uncorrelated from category to category. For each category, we then generated p/L patterns independently with a specified correlation coefficient c with the corresponding prototype. We show in Fig 7 the results of simulations with L = 5, f = 0.2 and ǫ = 3. The figure shows that the learning rule reaches a capacity that is essentially independent of c, in the range 0 ≤ c ≤ 0.75. Statistical properties of the connectivity matrix We next investigated the statistical properties of the connectivity matrix after the learning pro- cess. Previous studies have shown that the distribution of synaptic weights in perceptrons with excitatory synapses becomes at maximal capacity a delta function at zero weight, plus a truncated Gaussian for strictly positive weights [25, 28–30]. Our model differs from this setting because of the global inhibitory feedback. Despite this difference, the distribution of weights in our network bear similarities with the results obtained in these previous studies: the distribution exhibits a peak at zero weight (‘silent’, or ‘potential’ synapses), while the distribution of strictly positive weights resembles a truncated Gaussian. Finally, the fraction of silent synapses increases with the robustness parameter (see Fig 8). 13 C      2 1.5 1 0.5 y t i s n e D 0 0 (cid:1) = 0 P(w<0.05)=0.01 1 3 Weight strength 2 (cid:1) = 1.2 P(w<0.05)=0.250 (cid:1) = 0.66 P(w<0.05)=0.245 2 1.5 1 0.5 2 1.5 1 0.5 0 0 4 2 4 6 0 0 2 4 6 Figure 8: Synaptic weight distributions. Comparing the distributions of the synaptic weights at critical capacity for three different values of robustness obtained from simulation. The distri- bution of weights approaches a Dirac-delta distribution at zero plus a truncated Gaussian. As the patterns become more robust, the center of the partial Gaussian shifts towards the left, and the number of silent synapses increases. We have also computed the degree of symmetry of the weight matrix. The symmetry degree is computed as the Pearson correlation coefficient between the reciprocal weights in pairs of neurons. We observe a general trend towards an increasingly symmetric weight matrix as more patterns are stored, for all values of the robustness parameter ǫ (see Fig 9). 1 0.8 0.6 0.4 0.2 (cid:2) = 0 (cid:2) = 0.6 (cid:2) = 1.2 ) j i i w j, w ( r r o C 0 0 0.5 1 M   1.5 (cid:1)) 2 Figure 9: The degree of symmetry of the weight matrix. The Pearson correlation co- efficient between wij and wji is computed at different values of α for three values of ǫ. As α increases the weight matrix tends to be more symmetric, but gets saturated for high α. For the same values of α, as the robustness increases, the correlation also increases, so the weight matrix becomes more symmetric. Error bars (across 10 runs) are smaller than the symbols. 14 Discussion We presented a biologically-plausible learning rule that is characterized by three thresholds, and is able to store memory patterns close to the maximal storage capacity in a recurrent neural networks without the need of an explicit “error signal”. We demonstrated how the learning rule can be considered a transformed version of the PLR in the limit of a strong external field. Our network implements the separation between excitatory and inhibitory neurons, with learning occurring only at excitatory-to-excitatory synapses. We simulated a recurrent network with N = 1001 binary neurons, reaching to αc = 1.6 at zero basin size. We then used a robustness parameter ǫ to enlarge the basin size. The simulations showed that we are close to the theoretical capacity across the whole investigated range of values of ǫ. We expect that as N increases and the learning rate gets smaller, this difference would go to zero. Two crucial ingredients of the 3TLR are necessary: (1) strong external inputs, (2) three learning thresholds which are set according to the statistics of inputs to the neuron. The learning rule only uses information that is local to a synapse and corresponding neurons. Like classic Hebbian learning rules, our 3TLR works in an online fashion. In addition, it can also perform as a ‘palimpsest’ [31–33]: in case the total number of patterns exceeds the maximal capacity (at a certain basin size) the network begins to forget patterns that are not being presented anymore. Comparison with other learning rules The 3TLR can be framed in the setting of the classic Bienenstock-Cooper-Munro (BCM) the- ory [34, 35], with additional requirements to adapt it to the attractor network scenario. The original BCM theory uses firing-rate units, and prescribes that synaptic modifications should be proportional to (1) the synaptic input, and (2) a function φ (v) of the total input v (or, equiv- alently, of the total output). The function φ (v) is subject to two conditions: (1) φ (v) ≥ 0 (or ≤ 0) when v > θ (or < θ, respectively); (2) φ (0) = 0. The parameter θ is also assumed to change, but on a longer time scale (such that the changes reflect the statistics of the inputs); this (metaplastic) adaptation has the goal of avoiding the trivial situations in which all inputs elicit indistinguishable responses. This (loosely specified) framework ensures that, under reasonable conditions, the resulting units become highly selective to a subset of the inputs, and has been mainly used to model the developmental stages of primary sensory cortex. The arising selectiv- ity is spontaneous and completely unsupervised: in absence of further specifications, the units become selective to a random subset of the inputs (e.g. depending on random initial conditions). Our model is defined on simpler (binary) units; however, if we define φ (v) = Θ (v − θ) Θ (θ1 − v)− Θ (θ − v) Θ (v − θ0), then φ behaves according to the prescriptions of the BCM theory. Further- more, we have essentially assumed the same slow metaplastic adaptation mechanism of BCM, even though we have assigned this role explicitly to the inhibitory part of the network (see Ma- terials and Methods). On the other hand, our model has additional requirements: (1) φ (v) = 0 when v < θ0 or v > θ1, (2) plasticity occurs during presentation of external inputs, which in turn are strong enough to drive the network towards a desired state. The second requirement ensures that the network units become selective to a specific subset of the inputs, as opposed to a random subset as in the original BCM theory, and thus that they are able to collectively behave as an attractor network. The first requirement ensures that each unit operates close to critical capacity. Indeed, these additional requirements involve extra parameters with respect to the BCM theory, and we implicitly assume these parameters to also slowly adapt according to the statistics of the inputs during network formation and development. A variant of the BCM theory, known as ABS rule [36, 37] introduced a lower threshold for LTD, analogous to our θ0, motivated by experimental evidence; however, a high threshold for LTP, 15 analogous to our θ1, was not used there, or — to our knowledge — in any other BCM variant. The idea of stopping plasticity above some value of the ‘local field’ has been introduced previously to stabilize the learning process in feed-forward networks with discrete synapses [38–40]. Our study goes beyond these previous works in generalizing such a high threshold to recurrent networks, and showing that the resulting networks achieve close to maximal capacity. Comparison with data and experimental predictions In vitro experiments have characterized how synaptic plasticity depends on voltage [41] and fir- ing rate [42], both variables that are expected to have a monotonic relationship with the total excitatory synaptic inputs received by a neuron. In both cases, a low value of the controlling variable leads to no changes; intermediate values lead to depression; and high values to potenti- ation. These three regimes are consistent with the three regions for v < θ1 in Fig 2. The 3TLR predicts that a fourth region should occur at sufficiently high values of the voltage and/or firing rates. Most of the studies investigating the dependence of plasticity on firing rate or voltage have not reported a decrease in plasticity at high values of the controlling variables, but these studies might have not increased sufficiently such variables. To our knowledge, a single study has found that at high rates, the plasticity vs rate curve is a decreasing function of the input rate [43]. Another test of the model consists in comparing the statistics of the synaptic connectivity with experimental data. As it has been argued in several recent studies [25,28,30,44,45], networks with plastic excitatory synapses are generically sparse close to maximal capacity, with a connection probability that decreases with the robustness of information storage, consistent with short range cortical connectivity [46,47]. Our network is no exception, though the fraction of silent synapses that we observe is significantly lower than in models that lack inhibition. Furthermore, network that are close to maximal capacity tends to have a connectivity matrix that has a significant degree of symmetry, as illustrated by the over-representation of bidirectionally connected pairs of neurons, and the tendency of bidirectionally connected pairs to form stronger synapses than unidirectionally connected pairs as observed in cortex [47,48], except in barrel cortex [49]. Again, the 3TLR we have proposed here reproduces this feature (Fig 9), consistent with the fact that the rule approaches the optimal capacity. Future directions Our network uses the simplest possible single neuron model [50]. One obvious direction for future work would be to implement the learning rule in a network of more realistic neuron models such as firing rate models or spiking neuron models. Another potential direction would be to understand the biophysical mechanisms leading to the high threshold in the 3TLR. In any case, we believe the results discussed here provide a significant step in the quest for understanding how learning rules in cortical networks can optimize information storage capacity. Materials and Methods Simulation The main equations of the network, the neuron model, the learning rule, and the criteria for stopping the learning algorithm are outlined in the Results section, Eqs. 1-7. We present here additional details about network simulations. 16 Network setup before learning process Before applying the learning rule, we required the network to have stable dynamics around a desired activity level f . A network with only excitatory neurons is highly unstable and typically converges towards the trivial all-off and all-on states; therefore, we implemented a global inhibi- tion such that the network operates around activity level f . The basal inhibitory term (H0) and the inhibitory reaction term (H1) are defined as: H0 = (N − 1)(f ¯w − ψ) + H−1(f )p(N − 1)f σw H1 = f γ√N − 1 (11) (12) 2 erfc(cid:16) x√2(cid:17) and H−1 is the inverse of H, ψ is defined as θ = (N − 1)ψ; ¯w where H (x) = 1 and σw are the mean and standard deviation of the synaptic weights, respectively. With these definitions the network dynamics is stable in the sense that the activity level converges to f very fast, regardless of the initial condition. In Eq. 11, we see that H0 depends on the activity level f and on the standard deviation of the weights σw. In the dense regime, f = 0.5, we have H−1(0.5) = 0, therefore the rightmost term of Eq. 11 vanishes, which means that in this regime H0 is independent of σw. However, in sparser regimes, the network must be endowed with a mechanism to adjust for the changes in standard deviation, otherwise the learning process would bring the network out of the stable state, changing the basal activity level. In contrast, the mean synaptic efficacy ¯w does not change significantly during the learning process. In all our simulations, the initial values for {wij} were sampled from a Gaussian distribution with mean and standard deviation equal to one, after which negative values were set to zero. This has the effect the ¯winit is slightly higher than one. We also set wii = 0 for all i. ij Table 1 shows the values of the parameters used in the simulations, in the dense and sparse regimes. Table 1: Table of parameters in the simulation ij Parameter name N λ = ¯winit f ψ θ η γ # of interations (learning) # of trials in test phase Value in dense regime Value in sparse regime 1001 ≈1.08 0.5 0.35 350 0.01 [0.001 when ǫ = 0] 6.0 1000 [10000 when ǫ = 0] 50 1001 ≈1.08 0.2 0.35 350 0.01 [0.001 when ǫ = 0] 12.0 1000 [10000 when ǫ = 0] 50 Direct comparison between the 3TLR and the PLR In order to determine the degree to which the 3TLR is able to mimic the PRL, and the effect of deviations from the latter rule, we tested both rules on the same tasks. In these simulations, every part of the simulation code was kept identical — including the pseudo-random numbers used to choose the initial state and the arbitrary permutations for the update order of the units — except for the learning rule. We tested the network in the dense case f = 0.5, at ǫ = 3, varying 17 1 0.8 0.6 0.4 0.2 (cid:2) = 6) 1245 6 P4571245 6 (cid:2) = 12) 1.8 1.6 1.4 1.2 1 0.8 0.6 0.4 0.2 ) 6 = (cid:2) j# w i 0 0 0.05 0.1    0 0.2 0.15 (cid:1)) Figure 10: Direct comparions of the 3TLR and the PLR. Success probability for the 3TLR at γ = 6 (blue curve, left axis) and the PLR (red curve); the results for the 3TLR at γ = 12 are identical to those of the PLR (red curve). The orange points show the absolute difference of weights between the final values of the weights for the PLR at γ = 6 and the PLR (right axis): the points show the median of the distribution, while the error bars span the 5th-95th percentiles, showing that, while the distribution is concentrated at near-zero values, outliers appear at the critical capacity of the 3TLR algorithm. (Note that the average value of the weights is in all cases approximately 1.08; also compare the discrepancies with the overall distribution of the weights, Fig 8.) the storage load α, using 10 samples for each point. We compared the probability of solving the learning task and the distribution of the discrepancies (absolute value of the differences) in the values of the resulting synaptic weights. We tested two values of the parameter γ, 6 (as in Fig 4) and 12. We found that at γ = 12 there was absolutely no difference between the two rules, while at γ = 6 the 3TLR performed slightly worse, and significant deviations from the PLR started to appear close to the maximal capacity of the 3TLR (see Fig 10). Analytical calculation of the storage capacity at infinite N Entropy calculation In this section, we present the details of the calculations for the typical storage capacity of our network in the limit of N → ∞, using the Gardner analysis [20, 28]. typically be found. The capacity is defined as the maximum value of α = p/N such that a solution to Eq. 7 can We can rewrite Eq. 7 as ∀i : αN Yµ=1 where Θ (2ξµ i − 1)  wij ξµ N ξµ N Xj=1 Xj=1 j − f N i − H0 − λ   − θ H0 = N f ¯w − θ + H−1 (f ) σwpf N λ = ¯w  − f ǫ√N  = 1 (13) (14) (15) 18 S        p   !     " $ f f % & % n ' % ( % ) * % % n + L - . 3 / L - 0 Eq. 13 becomes: ∀i : αN Yµ=1 Θ (2ξµ i − 1)  N Xj=1 (wij − ¯w) ξµ i − H−1 (f ) σwpf N  − f ǫ√N  = 1 Let us now consider a single unit i. We write σµ i = (2ξµ as Wij = wij ¯w − 1 ∈ [−1,∞), and also define i − 1), and re-parametrize the weights (16) (17) (18) (19) T = H−1 (f )pf K = . ǫ ¯w Dropping the index i and neglecting terms of order 1, we obtain: αN Yµ=1 Θ σµ  Wjξµ j − T σw ¯w N Xj=1 √N  − f K√N  = 1 Our goal is to compute the quenched entropy of this problem, i.e. the scaled average of the logarithm of the volume of W which satisfies the above equation: S = = 1 N hlog V i{ξµ,σµ} N *logZ 1 Yj=1 N (dWjΘ (Wj + 1)) αN Yµ=1 Θ σµ  N Xj=1 Wj ξµ j − σw ¯w T√N  − f K√N + (20) {ξµ,σµ} The computation proceeds along the lines of [20, 28], by using the so-called replica trick to perform the average of the logarithm of V , exploiting the identity: hlog V i = lim n→0 hV ni − 1 n , (21) performing the computation for integer values of n and using an analytical continuation to perform the limit n → 0. we perform the calculation using the replica-symmetric (RS) Ansatz, which is believed to give exact results in the case of perceptron models with continuous weights. The final expression for the entropy depends on six order parameters; the first three are Q, q and M , whose meaning is Q = q = M = (Wj)2 W a j W b j Wj 1 1 N Xj N Xj √N Xj 1 where we used W a and W b to denote two different replicas of the system, which can simply be interpreted as two independent solutions to the constraint equation. Q is called the self-overlap, 19 are the conjugate quantities Q, q and M . The entropy expression is: in our case, while q is the mutual-overlap. The remaining order parameters ¯w (cid:1)2 and is equal to(cid:0) σw where q q S(cid:16)Q, q, M, Q, q, M(cid:17) = −(cid:18)Q Q − 2 (cid:19) + αZA (Q, q, M ) + ZW (cid:16) Q, q, M(cid:17) ZA (Q, q, M ) = Z Du*ln H K − σ(cid:0)M − T√Q(cid:1) + u (1 − f )√q ZW (cid:16) Q, q, M(cid:17) = Z Du ln(cid:18)Z ∞ 2 (q−2 Q)W 2+W (u√q− M)(cid:19) . (1 − f )√Q − q W e− 1 −1 !!+σ (22) (23) (24) 2 u x Du = 1 We used the usual notation Du ≡ du e− 2√2π 2 erfc(cid:16) x√2(cid:17). In the following, we will also use the shorthand G (x) = G(x) H (x) = R ∞ H(x) . We also used the notation h·iσ to denote the average over the output σ, i.e. hϕ (σ)iσ = f ϕ (1) + (1 − f ) ϕ (−1) for any function ϕ. The value of the order parameters is found by extremizing S. The notation and the following computations can be simplified using: = du G (u) to denote Gaussian integrals, and defined ∆Q = Q − q tσ (u) = K − σ(cid:0)M − T√Q(cid:1) + u (1 − f )√q (1 − f )√∆Q 2 ∆ QW 2+W (u√q− M) ∆ Q = q − 2 Q ν (u, W ) = e− 1 The extremization of S then results in the system of equations: ∆ Q = q = α p(Q − ∆Q) ∆QZ Du u hG (tσ (u))iσ ∆QZ Du hG (tσ (u)) tσ (u)iσ + ∆ Q α 0 = Z Du hG (tσ (u)) σiσ Q = Z Du R ∞ −1 dW W 2ν (u, W ) R ∞ −1 dW ν (u, W ) √q Z Du uR ∞ −1 dW W ν (u, W ) R ∞ −1 dW ν (u, W ) 0 = Z Du R ∞ −1 dW W ν (u, W ) R ∞ −1 dW ν (u, W ) 1 ∆Q = 20 (25) (26) (27) (28) (29) (30) (31) (32) (33) (34) The integrals over dW in the last three equations can be performed explicitly, yielding: Q = q + M 2 + ∆ Q ∆ Q2 ∆Q = 1 ∆ Q + 0 = − M ∆ Q Critical capacity + 3 1 + 1 ∆ Q 2 Z Du (cid:16)upq − M − ∆ Q(cid:17) G  − q∆ Qq Z Du uG  − u√q − M + ∆ Q q∆ QZ Du G  − q∆ Q q∆ Q u√q − M + ∆ Q     1 u√q − M + ∆ Q q∆ Q   (35) (36) (37) At critical capacity, the space of the solutions shrinks to a point, and the mutual overlap tends to become equal to the self overlap: q → Q, i.e. ∆Q → 0. In this limit, the conjugate order parameters diverge as: q = ∆ Q = M = C ∆Q2 A ∆Q B√C ∆Q (38) (39) (40) Using these conditions, and calling αc the critical value of α, the saddle point equations, 29 to 34, become: 1 A Q = A = H(cid:18)B − A(cid:16)C − B√C(cid:17) √C(cid:19) √C A (cid:18)G(cid:18)B − C = αcQ(cid:10)(cid:0)1 + τ 2 A = αc hH (τσ)iσ 0 = hσ (G (τσ) − τσH (τσ))iσ 0 = A √C(cid:19) − BA(cid:19) − (1 − A) σ(cid:1) H (τσ) − τσG (τσ)(cid:11)σ (41) (42) (43) (44) (45) (46) (47) where we defined τσ = σ(cid:0)M − T√Q(cid:1) − K (1 − f )√Q These equations can be solved numerically to find the six parameters αc, Q, A, B, C and M . Note that in the special case K = 0 these equations have a degenerate solution with Q = 0 and the same αc as in the case of unbounded synaptic weights (e.g. αc = 2 for f = 0.5). This is because in that case the original problem has the property that scaling all weights by a factor of x is equivalent to scaling the boundary ¯w by a factor of x−1 (see Eq. 16); therefore, the optimal strategy is to exploit this property by setting x → 0, i.e. effectively reducing the problem to the unbounded case. Of course, this strategy can only be pursued up to the available precision in a practical setting. 21 Acknowledgments References 1. Hopfield JJ. Neural networks and physical systems with emergent collective computational abilities. Proc Natl Acad Sci USA. 1982;79:2554–2558. 2. Amit DJ. Modeling brain function. Cambridge University Press; 1989. 3. Hertz J, Krogh A, Palmer RG. Introduction to the Theory of Neural Computation. Addison-Wesley, Redwood City; 1991. 4. Amit DJ, Brunel N. Model of global spontaneous activity and local structured activity during delay periods in the cerebral cortex. Cerebral Cortex. 1997;7:237–252. 5. Brunel N, Wang XJ. Effects of neuromodulation in a cortical network model of object working memory dominated by recurrent inhibition. J Comput Neurosci. 2001;11:63–85. 6. Mongillo G, Barak O, Tsodyks M. Synaptic Theory of Working Memory. Science. 2008;319:1543. 7. Barak O, Tsodyks M. Working models of working memory. Curr Opin Neurobiol. 2014;25:20–24. 8. Fuster JM, Jervey JP. Inferotemporal neurons distinguish and retain behaviourally relevant features of visual stimuli. Science. 1981;212:952–955. 9. Miyashita Y. Neuronal correlate of visual associative long-term memory in the primate temporal cortex. Nature. 1988;335:817–820. 10. Miyashita Y, Chang HS. Neuronal correlate of pictorial short-term memory in the primate temporal cortex. Nature. 1988;331:68–70. 11. Nakamura K, Kubota K. Mnemonic firing of neurons in the monkey temporal pole during a visual recognition memory task. J Neurophysiol. 1995;74:162–178. 12. Fuster JM, Alexander G. Neuron activity related to short-term memory. Science. 1971;173:652–654. 13. Funahashi S, Bruce CJ, Goldman-Rakic PS. Mnemonic coding of visual space in the monkey’s dorsolateral prefrontal cortex. J Neurophysiol. 1989;61:331–349. 14. Romo R, Brody CD, Hern´andez A, Lemus L. Neuronal correlates of parametric working memory in the prefrontal cortex. Nature. 1999;399:470–474. 15. Amit DJ, Gutfreund H, Sompolinsky H. Storing infinite numbers of patterns in a spin-glass model of neural networks. Phys Rev Lett. 1985;55:1530–1531. 16. Sompolinsky H. Neural networks with nonlinear synapses and a static noise. Phys Rev A. 1986;34:2571–2574. 17. Derrida B, Gardner E, Zippelius A. An exactly solvable asymmetric neural network model. Europhys Lett. 1987;4:167–173. 22 18. Tsodyks M, Feigel’man MV. The enhanced storage capacity in neural networks with low activity level. Europhys Lett. 1988;6:101–105. 19. Buhmann J, Divko R, Schulten K. Associative memory with high information content. Phys Rev A. 1989;39:2689–2692. 20. Gardner EJ. The phase space of interactions in neural network models. J Phys A: Math Gen. 1988;21:257–270. 21. Rosenblatt F. Principles of neurodynamics. Spartan Books, New York; 1962. 22. Marr D. A theory of cerebellar cortex. J Physiol. 1969;202:437–470. 23. Albus JS. A theory of cerebellar function. Mathematical Biosciences. 1971;10:26–51. 24. Ito M, Sakurai M, Tongroach P. Climbing fibre induced depression of both mossy fibre re- sponsiveness and glutamate sensitivity of cerebellar Purkinje cells. J Physiol. 1982;324:113– 134. 25. Clopath C, Nadal JP, Brunel N. Storage of correlated patterns in standard and bistable Purkinje cell models. PLoS Comput Biol. 2012;8:e1002448. 26. Parga N, Virasoro MA. The ultrametric organization of memories in a neural network. J Phys France. 1986;47:1857–1864. 27. Brunel N, Carusi F, Fusi S. Slow stochastic Hebbian learning of classes in recurrent neural networks. Network. 1998;9:123–152. 28. Brunel N, Hakim V, Isope P, Nadal JP, Barbour B. Optimal information storage and the distribution of synaptic weights: perceptron versus Purkinje cell. Neuron. 2004;43:745–57. 29. Brunel N, van Rossum MC. Lapicque’s 1907 paper: from frogs to integrate-and-fire. Biol Cybern. 2007;97:337–339. 30. Clopath C, Brunel N. Optimal properties of analog perceptrons with excitatory weights. PLoS Comput Biol. 2013;9:e1002919. 31. M M´ezard, Nadal JP, Toulouse G. Solvable models of working memories. J Physique. 1986;47:1457–. 32. Parisi G. A memory which forgets. J Phys A: Math Gen. 1986;19:L617. 33. Amit DJ, Fusi S. Dynamic learning in neural networks with material synapses. Neural Computation. 1994;6:957–982. 34. Bienenstock E, Cooper L, Munro P. Theory for the development of neuron selectivity: orientation specificity and binocular interaction in visual cortex. J Neurosci. 1982;2:32–48. 35. Jedlicka P. Synaptic plasticity, metaplasticity and BCM theory. Bratislavsk´e lek´arske listy. 2002;103(4/5):137–143. 36. Brocher S, Artola A, Singer W. Intracellular injection of Ca2+ chelators blocks induction of long-term depression in rat visual cortex. Proceedings of the National Academy of Sciences. 1992;89(1):123–127. 23 37. Artola A, Singer W. Long-term depression of excitatory synaptic transmission and its relationship to long-term potentiation. Trends in neurosciences. 1993;16(11):480–487. 38. Amit Y, Mascaro M. Attractor networks for shape recognition. Neural Comput. 2001;13:1415–1442. 39. Fusi S, Drew PJ, Abbott LF. Cascade models of synaptically stored memories. Neuron. 2005;45:599–611. 40. Brader JM, Senn W, Fusi S. Learning real-world stimuli in a neural network with spike- driven synaptic dynamics. Neural Comput. 2007;19:2881–2912. 41. Ngezahayo A, Schachner M, Artola A. Synaptic activity modulates the induction of bidi- rectional synaptic changes in adult mouse hippocampus. J Neurosci. 2000;20:2451–2458. 42. Kirkwood A, Rioult MC, Bear MF. Experience-dependent modification of synaptic plas- ticity in visual cortex. Nature. 1996;381:526–528. 43. Wang H, Wagner JJ. Priming-induced shift in synaptic plasticity in the rat hippocampus. J Neurophysiol. 1999;82:2024–2028. 44. Barbour B, Brunel N, Hakim V, Nadal JP. What can we learn from synaptic weight distributions? Trends Neurosci. 2007;30:622–629. 45. Chapeton J, Fares T, LaSota D, Stepanyants A. Efficient associative memory stor- age in cortical circuits of inhibitory and excitatory neurons. Proc Natl Acad Sci USA. 2012;109:E3614–3622. 46. Kalisman N, Silberberg G, Markram H. The neocortical microcircuit as a tabula rasa. Proc Natl Acad Sci U S A. 2005;102:880–885. 47. Song S, Sjostrom PJ, Reigl M, Nelson S, Chklovskii DB. Highly nonrandom features of synaptic connectivity in local cortical circuits. PLoS Biol. 2005;3:e68. 48. Wang Y, Markram H, Goodman PH, Berger TK, Ma J, Goldman-Rakic PS. Heterogeneity in the pyramidal network of the medial prefrontal cortex. Nat Neurosci. 2006;9:534–542. 49. Lefort S, Tomm C, Floyd Sarria JC, Petersen CC. The excitatory neuronal network of the C2 barrel column in mouse primary somatosensory cortex. Neuron. 2009;61:301–316. 50. McCulloch WS, Pitts WA. A logical calculus of the ideas immanent in nervous activity. Bull Math Biophys. 1943;5:115–133. 24
1212.0076
2
1212
2012-12-04T03:15:35
Short term synaptic depression improves information transfer in perceptual multistability
[ "q-bio.NC", "math.DS", "nlin.PS" ]
Competitive neural networks are often used to model the dynamics of perceptual bistability. Switching between percepts can occur through fluctuations and/or a slow adaptive process. Here, we analyze switching statistics in competitive networks with short term synaptic depression and noise. We start by analyzing a ring model that yields spatially structured solutions and complement this with a study of a space-free network whose populations are coupled with mutual inhibition. Dominance times arising from depression driven switching can be approximated using a separation of timescales in the ring and space-free model. For purely noise-driven switching, we use energy arguments to justify how dominance times are exponentially related to input strength. We also show that a combination of depression and noise generates realistic distributions of dominance times. Unimodal functions of dominance times are more easily differentiated from one another using Bayesian sampling, suggesting synaptic depression induced switching transfers more information about stimuli than noise-driven switching. Finally, we analyze a competitive network model of perceptual tristability, showing depression generates a memory of previous percepts based on the ordering of percepts.
q-bio.NC
q-bio
SHORT TERM SYNAPTIC DEPRESSION IMPROVES INFORMATION TRANSFER IN PERCEPTUAL MULTISTABILITY ZACHARY P. KILPATRICK Abstract. Competitive neural networks are often used to model the dynamics of perceptual bistability. Switching between percepts can occur through fluctuations and/or a slow adaptive pro- cess. Here, we analyze switching statistics in competitive networks with short term synaptic depres- sion and noise. We start by analyzing a ring model that yields spatially structured solutions and complement this with a study of a space-free network whose populations are coupled with mutual inhibition. Dominance times arising from depression driven switching can be approximated using a separation of timescales in the ring and space-free model. For purely noise-driven switching, we use energy arguments to justify how dominance times are exponentially related to input strength. We also show that a combination of depression and noise generates realistic distributions of domi- nance times. Unimodal functions of dominance times are more easily differentiated from one another using Bayesian sampling, suggesting synaptic depression induced switching transfers more informa- tion about stimuli than noise-driven switching. Finally, we analyze a competitive network model of perceptual tristability, showing depression generates a memory of previous percepts based on the ordering of percepts. Key words. binocular rivalry, neural field, ring model, bump attractor, short term depression 1. Introduction. Ambiguous sensory stimuli with two interpretations can pro- duce perceptual rivalry [5]. For example, two orthogonal gratings presented to either eye lead to perception switching between one grating and then the next repetitively, a common paradigm known as binocular rivalry [31]. Perceptual rivalry can also be triggered by a single stimulus with two interpretations, like the Necker cube [39]. One notable feature of the switching process in perceptual rivalry is its stochasticity -- a histogram of the dominance times of each percept spreads across a broad range [17]. Senses other than vision also exhibit perceptual rivalry. When two different odorants are presented to the two nostrils, a similar phenomenon occurs with olfaction, termed "binaral" rivalry [55]. Similar experiences have been evoked in the auditory [13, 41] and tactile [10] system. Multiple principles govern the relationship between the strength of percepts and the mean switching statistics in perceptual rivalry [32]. "Levelt's propositions" relate stimulus contrast to the mean dominance times: (i) increasing the contrast of one stimulus increases the proportion of time that stimulus is dominant; (ii) increasing the contrast of one stimulus does not affect its average dominance time; (iii) increas- ing contrast of one stimulus increases the rivalry alternation rate; and (iv) increasing the contrast of both stimuli increases the rivalry alternation rate. There are also relationships between the properties of the input and the stochastic variation in the dominance times [6]. For instance, the distribution of dominance times is well fit by a gamma distribution [17, 30, 47]. The fact that dominance times are not expo- nentially distributed suggests some background slow adaptive process plays a role in providing a nonzero peak in the dominance histograms [44]. Two commonly proposed mechanisms for this adaptation are spike frequency adaptation and short term synap- tic depression [29, 49, 43]. An even stronger case can be made of the existence of adaptation in perceptual processing networks by examining results of experiments on perceptual tristability [21]. Here, perception alternates between three possible choices and subsequent switches are determined by the previous switch [38]. This memory suggests switches in perceptual multistability are not purely noise-driven [37]. Most theoretical models of perceptual rivalry employ two pools of neurons, each 1 selective to one percept, coupled to one another by mutual inhibition [33, 29, 43, 42]. With no other mechanisms at work, such architectures lead to winner-take-all states, where one pool of neurons inhibits the other indefinitely [48]. However, switches between the dominance of one pool and the other can be initiated with the inclusion of fluctuations [37] or an adaptive process [29, 43]. Combining the two effects leads to dominance times that are distributed according to the gamma distribution [29, 44, 47]. Slow adaptation and noise thus serve as agents for the sampling of the stimulus through network activity. A mutual inhibitory network would otherwise remain in the winner-take-all state indefinitely. In light of these observations, we wish to consider the role adaptive mechanisms play in properly sampling ambiguous stimuli in the context of a mutual inhibitory network. Purely fluctuation driven switching would provide a noisy sample of the two percepts, but pure adaptation would provide an extremely reliable sampling of percept contrast [44]. Thus, as the level of adaptation is increased and noise is decreased, one would expect that the ability of mutual inhibitory networks to encode information about ambiguous stimuli is vastly improved. A major point is that it is also vastly improved over networks without any adaptation at all. We focus specifically on short term synaptic depression [1, 45]. Using parametrized models, we will explore how synaptic depression improves the ability of a network to extract stimulus contrasts. First, we will be concerned with how much information can be determined about the contrast of each of the two percepts of an ambiguous stimulus. In the case of a winner-take-all solution, only information about a single percept could be known, since the pool of neurons encoding the other percept would be quiescent. We will study this problem using an anatomically moti- vated neural field model of an orientation column with synaptic depression [54, 25], given by (1). We find that increasing the strength of synaptic depression from zero leads to a bifurcation whereby rivalrous oscillations onset. When rivalrous switching occurs through a combination of depression and noise, we show stronger depression improves the transfer of information using simple Bayesian inference [24]. We also analyze a competitive network model with depression and noise (6) to help study the combined effects of noise and depression on perceptual switching. In particular, we will show that the presence of synaptic depression increases the information relayed by the output of the network. Finally, we will show trimodal stimuli to a neural field model with synaptic depression can generate oscillations where each mode spends time in dominance. To deeply analyze the relative contributions of noise and depression to this switching process, we study a reduced model. This reveals depression generates a history dependence in switching that would not arise in the network with purely noise-driven switching. 2. Materials and Methods. 2.1. Ring model with synaptic depression. As a starting point, we will consider a model for processing the orientation of visual stimuli [3, 7] which also includes short term synaptic depression [54, 25]. Since GABAergic inhibition is much faster than AMPA-mediated excitation [23], we make the assumption that inhibition is slaved to excitation as in [2]. Reduction this disynaptic pathway and assuming depression acts on on excitation [45], we then have the model [54, 26] (cid:90) π/2 −π/2 τm ∂u(x, t) ∂t = −u(x, t) + w(x − y)q(y, t)f (u(y, t))dy + I(x) + ξ(x, t), (1a) 2 τ ∂q(x, t) ∂t = 1 − q(x, t) − βq(x, t)f (u(x, t)). (1b) Here u(x, t) measures the synaptic input to the neural population with stimulus pref- erence x at time t, evolving on the timescale τm. Firing rates are given by taking the gain function f (u) of the synaptic input, which we usually proscribe to be [51] and often take the high gain limit γ → ∞ for analytical convenience, so [2, 26] f (u) = H(u − κ) = : u < κ, : u ≥ κ. f (u) = 1 + e−γ(u−κ) , 1 (cid:26) 0 1 (2) (3) External input, representing flow from upstream in the visual system is prescribed by the time-independent function I(x) [3, 7]. For the majority of our study of (1), we employ the bimodal stimulus I(x) = −I0 cos(4x) + Ia sin(2x), (4) representing stimuli at the two orthogonal angles −π/4 and π/4 and I0 controls the mean of each peak and Ia controls the level of asymmetry between the peaks. Ef- fects of noise are described by the stochastic processe (cid:104)ξ(x, t)(cid:105) with (cid:104)ξ(x, t)(cid:105) = 0 and (cid:104)ξ(x, t)ξ(y, s)(cid:105) = C(x − y)δ(t − s). For simplicity, we assume the spatial correlations have a cosine profile C(x) = π cos(x). Synaptic interactions are described by the inte- gral term, so w(x − y) describes the strength (amplitude of w) and net polarity (sign of w) of synaptic interactions from neurons with stimulus preference y to those with preference x. Following previous studies, we presume the modulation of the synaptic strength is given by the cosine w(x − y) = cos(2(x − y)), (5) so neurons with similar orientation preference excite one another and those with dissimilar orientation preference disynaptically inhibit one another [3, 15]. The factor q(x, t) measures of the fraction of available presynaptic resources, which are depleted at a rate βf [1, 45], and are recovered on a timescale specified by the time constant τ [11]. By setting τm = 1, we can assume time evolves on units of the excitatory synaptic time constant, which we presume to be roughly 10ms [18]. Experimental observations have shown synaptic resources specified q are recovered on a timescale of 200-800ms [1, 46], so we require τ is between 20 and 80, usually setting it to be τ = 50. Our parameter β can then be varied independently to adjust the effective depletion rate of synaptic depression. 2.2. Idealized competitive neural network. We also study space-free com- petitive neural networks with synaptic depression [43]. In this way, we can make more progress analyzing switching behavior. As a general model of networks connected by mutual inhibition, we consider the system [29, 37, 43] uR(t) = −u1(t) + f (IR − qL(t)uL(t)) + ξ1(t), uL(t) = −u2(t) + f (IL − qR(t)uR(t)) + ξ2(t), τ qR(t) = 1 − qR(t) − βuR(t)qR(t), τ qL(t) = 1 − qL(t) − βuL(t)qL(t), 3 (6a) (6b) (6c) (6d) where uj represents the firing rate of the j = 1, 2 population. The strength of recurrent synaptic excitation within a population is specified by the parameter α, whereas the strength of cross-inhibition between populations is specified by β. Fluctuations are introduced into population j with the independent white noise processes ξj with (cid:104)xj(t)(cid:105) = 0 and (cid:104)ξj(t)ξj(s)(cid:105) = εδ(t − s). 2.3. Numerical simulation of stochastic differential equations. The spa- tially extended model (1) was simulated using an Euler-Maruyama method with a timestep dt = 10−4, using Riemann integration on the convolution term with N = 2000 spatial grid points. The space clamped competitive network (6) was also simulated using Euler-Maruyama with a timestep dt = 10−6. To generate histograms of dominance times, we simulated systems for 20000s (2 × 106 time units). 2.4. Fitting dominance time distributions. To generate the theoretical curves presented for exponentially distributed dominance times, we simply take the mean of dominance times and use it as the scaling in the exponential (40). For those densi- ties that we presume are gamma distributed, we solve a linear regression problem. Specifically, we look for the constants c1, c2, and c3 of f (T ) = ec1T c2 e−c3T (7) an alternate form of (42). Upon taking the logarithm of (7), we have the linear sum ln f (T ) = c1 + c2 ln T − c3T. (8) Then, we select three values of the numerically generated distribution pn(T n) along with its associated dominance times: (T n j ). We always choose T n 2 /2. It is then straightforward to solve the linear system 2 , pn 2 = arg maxT pn(T ) as well as T n 3 , pn 2 /2 and T n 2 ); (T n 1 = T n 3 ) where pn 1 , pn 1 ); (T n j = pn(T n 3 = 3T n  1 1 ln pn 2 ln pn 3 for the associated constants using the \ command in MATLAB. c2 c3 2 ln T n 1 ln T n 1 ln T n 1 1 −T n 2 −T n 3 −T n 3  c1  =  ln pn  3. Results. We now discuss several results that reveal the importance of synap- tic depression in transferring information about stimuli to competitive networks. This is initially shown by analyzing the ring model with depression (1). However, to carry out detailed analysis on stochastic switching a competitive network with depression, we must reduce (1) as well as analyzing an analogous model without space (6). 3.1. Deterministic switching in the ring model. To start we will consider the ring model with depression (1) in the absence of noise. We then have the deter- ministic system [54, 26, 27] (cid:90) π/2 ut(x, t) = −u(x, t) + τ qt(x, t) = 1 − q(x, t) − βq(x, t)f (u(x, t)). −π/2 w(x − y)q(y, t)f (u(y, t))dy + I(x), (9a) (9b) In previous work, versions of (9) have been analyzed to explore how synaptic de- pression can generate traveling pulses [54, 26], self-sustained oscillations [26], and spiral waves in two-dimensions [27]. Here, we will extend previous work that explored 4 Fig. 1. Three possible active states of the noise-free stimulus driven ring model with depression (9). (A) Winner take all (I0 = 0.6); (B) Rivalrous oscillations (I0 = 0.84); (C) Fusion (I0 = 1). Other parameters are θ = 0.5, β = 1, and τ = 50. input-driven oscillations in two-layer networks like (9) that possessed many statistics matching binocular rivalry [25]. We will think of (9) as a model of monocular ri- valry, since oscillations can be due to competition between representations in a single orientation column [3]. Competition between ocular dominance columns [25] is not necessary for our theory. For the purpose of exposition, we will employ specific forms for the functions of (9): cosine weight (5); a Heaviside firing rate function (3); and a bimodal input (4). Winner take all state. We start by looking for winner-take-all solutions to the deterministic system (9), such as that shown in Fig. 1(A). These states consist of a single activity bump arising in the network, representing only one of the two percepts contained in the bimodal stimulus (4). These are stationary in time, so ut = qt = 0, implying that u = U (x) and q = Q(x). Also, they are single bump solutions, so there will be a single region in space that is superthreshold (U (x) > κ). We assume the right stimulus is represented by a bump, although we can derive analogous results when the left stimulus is represented. We then have a steady state solution determined by (cid:90) π/4+a U (x) = Q(x) = [1 + βH(U (x) − κ)] π/4−a −1 , cos(2(x − y))Q(y)dy − I0 cos(4x) + Ia sin(2x), (10) (11) since U (x) > κ when x ∈ (π/4 − a, π/4 + a), so that by plugging (11) into (10) and using the trigonometric identity cos(2(x − y)) = cos(2x) cos(2y) + sin(2x) sin(2y) we have U (x) = A cos(2x) + B sin(2x) − I0 cos(4x) + Ia sin(2x), (12) where the multiplicative constants A, B can be computed (cid:90) π/4+a (cid:90) π/4+a 1 A = 1 + β π/4−a cos(2x)dx = 0, B = 1 1 + β π/4−a sin(2x)dx = sin(2a) 1 + β . Therefore, by simplifying the threshold condition, U (π/4 ± a) = κ, we have U (π/4 ± a) = sin(4a) 2(1 + β) + I0 cos(4a) + Ia cos(2a) = κ. (13) 5 The implicit equation (13) can be solved numerically using root finding algorithms. For symmetric inputs (Ia = 0), we can solve (13) explicitly (cid:34) 1 ±(cid:112)1 + 4(1 + β)2(I 2 2(1 + β)(I0 + κ) (cid:35) 0 − κ2) a = tan−1 1 2 . (14) Thus, we can fully characterize winner-take-all solutions U (x) = sin(2a) 1 + β sin(2x) − I0 cos(4x) + Ia sin(2x). (15) The advantage of having this solution is that we can relate the parameters of the model to the existence of the winner-take-all state, where we would expect to only see single bump solutions. To do so, we need to look at a second condition that must be satisfied, U (x) < κ for all x /∈ (π/4 − a, π/4 + a). Since the function (15) is bimodal across (−π/2, π/2), we check the other possible local maximum at x = −π/4 as U (π/4) = I0 − Ia − sin(2a) 1 + β < κ. (16) At the point in parameter space where the (16) is violated, a bifurcation occurs, so the winner-take-all state ceases to exist. This surface in parameter space is given by the equation I0 = κ + Ia + sin(2a) 1 + β , (17) along with the explicit formula for the bump half-width (14). Beyond the bifurcation boundary (17), one of two behaviors can occur. Either there is a symmetric two-bump solution that exists, the fusion state [52, 4, 43], or rivalrous oscillations [32, 5]. Fusion state. It has been observed in many experiments on ambiguous stimuli that sufficiently strong contrast rivalrous stimuli can be perceived as a single fused im- age [4, 8]. This should not be surprising, considering stereoscopic vision and audition behave in exactly this way [52]. However, the contrast necessary to evoke this state with dissimilar images is much higher than with similar images [5]. In the network (9), the fusion state (Fig. 1(C)) is represented as two disjoint bumps. Therefore (cid:34)(cid:90) −π/4+b (cid:90) π/4+a (cid:35) U (x) = 1 1 + β −π/4−b π/4−a + cos(2(x − y))dy − I0 cos(4x) + Ia sin(2x). Computing the integral terms, we find U (x) = S(x, a) − S(x, b) 1 + β − I0 cos(4x) + Ia sin(2x), (18) where S(x, y) = sin2(x + y) − sin2(x − y). The solution can be specified by requiring the threshold conditions U (−π/4 ± b) = U (π/4 ± a) = κ are satisfied cos(2a)[sin(2a) − sin(2b)] 1 + β cos(2b)[sin(2b) − sin(2a)] 1 + β + I0 cos(4a) + Ia cos(2a) = κ, + I0 cos(4b) − Ia cos(2b) = κ, (19) (20) 6 which we can solve numerically to relate the asymmetry of inputs Ia to the half- widths a, b of each bumps. In the the case of symmetric inputs, U (x) = −I0 cos(4x), and it is straightforward to find the two bump widths explicitly. We will now study rivalrous oscillations by simply constructing them using a fast-slow analysis. We can also numerically identify the boundary of various behaviors of (1), as shown in Fig. 3. Rivalrous oscillations. Oscillations can occur, where the two bump locations trade dominance successively (Fig. 1(B)). As in Levelt's proposition (i), increasing the contrast of a stimulus leads to that stimulus being in dominance longer. This infor- mation is not revealed when the system is stuck in a winner-take-all state. Therefore, the introduction of synaptic depression into the system (9) leads to an increase in information transfer. We will also examine how well (9) recapitulates Levelt's other propositions concerning the mean dominance of percepts. To analyze (9) for oscillations, we assume that the timescale of synaptic depression τ (cid:29) 1, long enough that we can decompose (9) into a fast and slow system [29, 25]. Synaptic input u then tracks the slowly varying state of the synaptic scaling term q. We also assume that q is essentially piecewise constant in space, in the case of the Heaviside nonlinearity (3), which yields u(x, t) ≈ cos(2(x − y))q(y, t)H(u(y, t) − κ)dy − I0 cos(4x), (21) (cid:90) π/2 −π/2 and q is governed by (9b). To start, we will also assume a symmetric bimodal input. This way, we can simply track q in the interior of one of the bumps, given qi(t) = q(π/4, t). Assuming a switch has just occurred, where the left bump has escaped suppression, to pin down the right, so τ qi(t) = 1 − qi(t), t ∈ (0, T ), qi(0) = q0, (22) where T is the amount of time each percept is in dominance, and q0 is the synaptic strength within a bump region immediately prior to its shutting off. After the right bump escapes dominance of the left bump τ qi(t) = 1 − qi(t) − βqi(t), t ∈ (T, 2T ). (23) Solving (22) and (23) simultaneously, we have which can be solved explicitly for the dominance time q0 = 1 + β e−T /τ − (1 − q0)e−2T /τ , 1 + β 1 + β (cid:34) β +(cid:112)β2 − 4(1 + β)(1 − q0)[(1 + β)q0 − 1] T = τ ln 2(1 + β)q0 − 2 (cid:35) , (24) so that we now must specify the value q0. We can examine the fast equation (21), solving for the form of the slowly narrowing right bump during its dominance phase u(x, t) = qi(t) = qi(t)(cid:2)sin2(x + a(t)) − sin2(x − a(t))(cid:3) − I0 cos(4x). cos(2(x − y))dy − I0 cos(4x) π/4−a(t) (25) (cid:90) π/4+a(t) 7 We can solve for the slowly evolving width a(t) of this bump by requiring the threshold condition u(π/4 ± a(t), t) = κ to yield qi(t) 2 sin(4a(t)) + I0 cos(4a(t)) = κ, and then solving using trigonometric identities (cid:34) qi(t) +(cid:112)qi(t)2 + 4(I 2 0 − κ2) (cid:35) . (26) 2(I0 + κ) a(t) = tan−1 1 2 We can also identify the maximal value of qi(t) = q0 which still leads to the right bump suppressing the left. Once qi(t) falls below q0, the other bump escapes suppression, flipping the dominance of the current bump. This is the point at which the other bump of (25) rises above threshold, as defined by the equation u(−π/4, t) = I0 − q0 sin(2a0) = κ. Combining this with the equation (26), we have an algebraic equation for q0 given (I0 − κ)2 = q2 0 2q2 8I 2 0 + 4(I 2 0 + 8I0κ + 2q2 0 + 4(I 2 0 + 4(I 2 0 − κ2) 0 − κ2) 0 − κ2) + 2q0 0 + 2q0 (cid:112)q2 (cid:112)q2 (cid:112)(I0 − κ)(3I0 + κ) 3I0 + κ , (27) which is straightforward to solve for 2I0 q0 = and we have excluded an extraneous negative solution. Interestingly, the amplitude of synaptic depression is excluded from (27), but we do know based on (22) and (23) that q0 ∈ ([1 + β]−1, 1). This establishes a bounded region of parameter space in which we can expect to find rivalrous oscillations, which we use to construct a partitioning of parameter space in Fig. 3. We can also now approximate the dominance time using (24) with (27), as shown in Fig. 2(D). In the case of an asymmetric bimodal input (Ia > 0), we can also solve for explicit approximations to the dominance times of the right TR and left TL populations. Following the same formalism as for the symmetric input case (cid:34) (cid:34) β + qd +(cid:112)(β + qd)2 − 4(1 + β)(1 − qL)[(1 + β)qR − 1] β − qd +(cid:112)(β − qd)2 − 4(1 + β)(1 − qR)[(1 + β)qL − 1] 2(1 + β)qR − 2 (cid:35) (cid:35) , , (28) (29) TR = τ ln TL = τ ln 2(1 + β)qL − 2 where qd = (1 + β)(qR − qL), in terms of the local values qL and qR of the synaptic scaling in the right and left bump immediately prior to their suppression. Notice in the case qL = qR, then qd = 0 and (28) and (29) both reduce to (24). We now need to examine the fast equation (21) to identify these two values. This is done by generating two implicit equations for the half-width aR and qR at the time of a switch qR 2 sin(4aR) + I0 cos(4aR) + Ia cos(2aR) = κ, I0 − Ia − qR sin(2aR) = κ, (30) (31) 8 Fig. 2. Dependence of rivalry on the amplitudes of the bimodal input (4). (A) Dominance times are both T ≈ 1s when input is symmetric (I0 = 0.84, Ia = 0). (B) Dominance time of right input (TR ≈ 0.9s) is longer than left (TL ≈ 0.6s) for asymmetric input (IR = 0.9, IL = 0.84). Notice TR is unchanged from case (A). (C) Dominance times are both shorter of higher contrast symmetric stimulus (I0 = 0.9, Ia = 0). (D) Increasing the strength of the symmetric (Ia = 0) bimodal input (4) decreases the dominance time T of both populations. Our theory (black) computed from fast-slow analysis (24) fits results of numerical simulations (blue) well. (E) For asymmetric inputs (Ia (cid:54)= 0), we find that varying IR = I0 + Ia while keeping IL = I0 − Ia fixed changes the dominance times of the left percept TL (blue) much more than that of the right percept TR (red). Other parameters are κ = 0.5, β = 1, and τ = 50. which we can solve explicitly for aR = cos−1 1 2 (cid:20) κ 2I0 (cid:21) , + 1 2 and (cid:112)(3I0 + κ)(I0 − κ) 2I0(IL − κ) , qR = (32) (33) where IL = I0 − Ia is the strength of input to the left side of the network. Likewise, we can find the value of the synaptic scaling in the left bump immediately prior to its suppression , (34) (cid:112)(3I0 + κ)(I0 − κ) 2I0(IR − κ) qL = where IR = I0 + Ia is the strength of input to the right side of the network. Using the expressions (33) and (34) we can now compute the dominance time formulae (28) and (29), showing the relationship between inputs and dominance times in Fig. 2 (E). Notice that all of Levelt's propositions are essentially satisfied. Changing the strength of the right stimulus IR has a very weak effect on the dominance time of the 9 Fig. 3. Partitions of parameter space into various stimulus induced states of (9). (A) Plotted as a function of network threshold κ and strength I0 of the bimodal input (4) when β = 1. (B) Plotted as a function of synaptic depression strength β and strength I0 of bimodal input (4) when κ = 0.5. Parameter τ = 50. right percept. However, it does increase the overall alternation rate and decrease the proportion of time the left percept remains in dominance. We also note there is a critical strength of synaptic depression β below which rivalrous oscillations do not occur. When synaptic depression is sufficiently strong, the winner-take-all state ceases to exist. Beyond this critical synaptic depression strength, the network either supports rivalrous oscillations or a fusion state. Either way, information is conveyed to the network that would otherwise be kept hidden. We show this in Fig. 3(B). In this way, synaptic depression can improve the information transfer of the network (9). In fact, we will show that it does so in a way that is much more reliable than noise. 3.2. Purely stochastic switching in the ring model. We will now study rivalrous switching brought about by fluctuations. In particular, we ignore depression and examine the noisy system (cid:35) (cid:34) (cid:90) π/2 −π/2 du(x, t) = −u(x, t) + w(x − y)f (u(y, t))dy + I(x) dt + dξ(x, t). (35) where (cid:104)ξ(x, t)(cid:105) = 0 and (cid:104)ξ(x, t)ξ(y, s)(cid:105) = εC(x− y)δ(t− s) defines the spatiotemporal correlations of the system. To start, we can simply consider (35) in the absence of noise ut(x, t) = −u(x, t) + w(x − y)f (u(y, t))dy + I(x). (36) (cid:90) π/2 −π/2 This model has been studied extensively [2, 3], so we will not perform an in depth analysis of stationary bump solutions. We are interested in the winner-take-all state. For a cosine weight (5), Heaviside firing rate (3), and bimodal stimulus (4) these can be computed as a special case of (15) and lie at either x = ±π/4, so U±(x) = sin(2a±) sin(2x) − I0 cos(4x) ± Ia sin(2x), (37) 10 Fig. 4. Purely noise induced switching of dominance in the depression-free ring model (35) (A) Numerical simulations of the system for the various input strengths of the symmetric (Ia = 0) bimodal input (4) with (A) I0 = 0.8, (B) I0 = 0.9, and (C) I0 = 1.0. Distributions of dominance times computed numerically (blue bars) with the exponential distribution (40) with numerically com- puted mean (cid:104)T(cid:105) (red) superimposed for (D) I0 = 0.8 has (cid:104)T(cid:105) ≈ 1.2s, (e) I0 = 0.9 has (cid:104)T(cid:105) ≈ 0.70s, and (f ) I0 = 1.0 has (cid:104)T(cid:105) ≈ 0.45s. Other parameters are κ = 0.5 and ε = 0.04. and we can apply the threshold condition U±(±π/4 + a±) = κ, so sin(4a±) + I0 cos(4a±) ± Ia cos(2a±) = κ. 1 2 In the case of a symmetric input, a± = a and we can solve (38) explicitly (cid:34) 1 +(cid:112)1 + 4(I 2 2(I0 + κ) (cid:35) . 0 − κ2) a = tan−1 1 2 (38) (39) Since we have no synaptic depression in the model (36), we cannot rely on deter- ministic mechanisms to generate switches between one winner-take-all state and an- other. Therefore, we will consider the effects of introducing a small amount of noise (0 < ε (cid:28) 1), reflective of synaptic fluctuations. We focus on the spatial correlation function C(x) = cos(x). Noise can generate switches in between the two dominant states (Fig. 4). As the strength of both input contrasts are increased, switches be- tween percepts occur more often. In fact, there is an exponential dependence of the mean dominance time (cid:104)T(cid:105) on the strength I0 of the bimodal input (4). We will pro- vide an argument, using energy methods, as to why this occurs in our analysis of the simplified system (6). We will study switching in the case that the suppressed bump comes on and shuts off the bump that is currently on. This mechanism is known as escape [48]. As shown in Fig. 4(A-C), dominance times decrease with input strength on average. Here escape is a noise induced effect [37], rather than a deterministic depression- induced effect [43, 25]. However, as opposed to depression-induced switching, there is a substantial spread in the possible dominance times for a given set of parameters (Fig. 4(D-F)). Thus, by examining two dominance times back to back, an observer would 11 Fig. 5. Sampling of the strength of each input based on noise-induced transitions. (A) Mean dominance time (cid:104)T(cid:105) as a function of the strength I0 of the symmetric (Ia = 0) bimodal input (4). (B) Probability that the right input IR was higher than the left output IL, based on the sampling n cycles (2n switches between percepts), in the case of symmetric inputs IL = IR = 0.9. Notice it takes close to 2000 cycles before p[IR > ILT ∗(n)] ≈ 0.5. Other parameters are κ = 0.5 and ε = 0.04. have difficulty telling if the input strengths were roughly the same or not. Recently, psychophysical experiments have been carried out where an observer must identify the higher contrast of two rivalrous percepts [36], showing humans perform quite well in at this task. Results of [36] suggest humans most likely use Bayesian inference in discerning information about visual percepts. This is in keeping with previous observations that humans' visual perception of objects likely carries out Bayesian inference [24]. Notice in Fig. 5(B) that the likelihood an observer assigns to IR > IL approaches 1/2 as the number of observations n increases. We compute p[IR > ILT ∗(n)], the probability an observer would presume IR > IR conditioned on dominance time pairs from n cycles T ∗(n) = , numerically here. How- ever, as the number of cycles n → ∞, the exponential distributions approximately defining the identical probability densities pR(TR) = pL(TL) = p(T ) will be fully sampled. We can calculate L , ..., T (n) T (1) R , T (1) L , T (2) R , T (2) R , T (n) L (cid:110) (cid:111) p(IR > ILT ∗(∞)) = (cid:90) ∞ (cid:90) x 0 0 p(x)p(y)dydx = 1 (cid:104)T(cid:105)2 e−(x+y)/(cid:104)T(cid:105)dydx = 1 2 , (cid:90) ∞ (cid:90) x 0 0 as in Fig. 5(B). In the case where depression, in the absence of noise, drives switches, we would expect this limit to be approached much more quickly. We will also examine how a combination of depression and noise affects an observer's ability to discern contrast differences. We explore this further in the case of asymmetric inputs, showing that dominance times are still specified by roughly exponential distributions as shown in Fig. 6. When IR > IL, even though the means satisfy (cid:104)TR(cid:105) > (cid:104)TL(cid:105) (Fig. 6(E)), the exponential distributions p(TR) and p(TL) have considerable variance, also given by the means. Therefore, randomly sampled values from these distributions may satisfy TR < TL. Were an observer to use one such sample as a means for guessing the inputs that generated them, they would guess IR < IL, rather than the correct IR > IL. In terms of conditional probabilities, we can expect situations where p(IR > ILT ∗(n)) < 1/2 12 Fig. 6. (A) Single realization of the stochastic neural field (35) with asymmetric (Ia > 0) (B) inputs IR = 0.92 and IL = 0.88, leads to longer dominance times for right percept TR. Expected likelihood p[IR > IRT ∗(n)] that the right input IR is stronger than left IL based on n comparisons of dominance times TR and TL sampled. Blue line is theoretical prediction (41) of the limit as n → ∞. Numerically computed dominance time distributions (blue bars) are well fit by the exponential distribution (40) for the (C) left ((cid:104)TL(cid:105) ≈ 0.5s) and (D) right ((cid:104)TR(cid:105) ≈ 1s) percepts. (E)Dependence of mean dominance times (cid:104)TR(cid:105) and (cid:104)TL(cid:105) on the strength of the right input IR. Black curves are best fits to exponential functions of IR. (F) Expected likelihood p[IR > ILT ∗(∞)] right input IR is stronger than left IL in the limit of high sample number n → ∞, as computed theoretically by (41). Other parameters are κ = 0.5, and ε = 0.04. for finite n, even though IR > IL. We can quantify this effect numerically, as shown in Fig. 6(F). In the limit n → ∞, we find uncertainty continues to creep in, since fluctuations continually give an observer misleading information. Since the marginal distributions are approximately exponential pj(Tj) = e−Tj /(cid:104)Tj(cid:105)/(cid:104)Tj(cid:105) j = L, R, (40) 13 Fig. 7. Switching in the stochastic ring model with depression (1) with asymmetric inputs (Ia > 0). (A) Single realization for asymmetric inputs with IR = 0.92 and IL = 0.88, which leads to right percept dominating longer. (B) Distribution of left percept dominances times pL(TL) over 1000s is well fit by a gamma distribution (42). (C) Distribution of right percept dominance times pR(TR) across 1000s is well fit by a gamma distribution (42). Other parameters are κ = 0.5, β = 0.2, τ = 50, and ε = 0.01. we can approximate the conditional probability p[IR > ILT ∗(∞)] = = (cid:90) ∞ (cid:90) x 0 0 1 (cid:104)TR(cid:105)(cid:105)TL(cid:105) pR(x)pL(y)dydx (cid:90) ∞ (cid:90) x e−x/(cid:105)TR(cid:105)e−y/(cid:105)TL(cid:105)dydx = 1 − 0 0 (cid:104)TL(cid:105) (cid:104)TR(cid:105) + (cid:104)TL(cid:105) = (cid:104)TR(cid:105) (cid:104)TR(cid:105) + (cid:104)TL(cid:105) . (41) Observe that the approximation we make using the formula (41) accurately estimates the limit p(IR > ILT ∗(∞)) as shown in Fig. 6(B). This is the likelihood that an observer performing Bayesian sampling of the probability densities (40) will predict IR > IL. Recent psychophysical experiments suggests humans would perform this task of contrast differentiation in this way [24, 36]. We see from our analysis that when switches are generated by noise, rather than deterministic depression, the means dominance times still obey Levelt's propositions to some extent (Fig. 6(E)). This would allow for an accurate comparison of input strengths IR and IL based on the means (cid:104)TR(cid:105) and (cid:104)TL(cid:105). However, when consider- ing a more realistic observer, that could only compare successive dominance times, accurately discerning the comparison of the input contrasts is more difficult. This be- comes much more noticeable when the input contrasts are quite close to one another, as in Fig. 6(F). We will explore now how introducing depression along with noise improves discernment of the input contrasts by an observer using simple comparison of dominance times. 3.3. Switching through combined depression and noise. We now study the effects of combining noise and depression in the full ring model of perceptual rivalry (1). Numerical simulations of (1) reveal that noise-induced switches occur robustly, even in parameter regimes where the noise-free system supports no rivalrous oscillations, as shown in Fig. 7. Rather than dominance times being distributed exponentially, they roughly follow a gamma distribution [17, 30] pj(Tj) = 1 σkΓ(k) j exp [−Tj/σ] , T k k > 1. (42) 14 Fig. 8. Comparing the probability densities of dominance times in the stochastic ring model with depression (1) for various levels of noise and depression. (A) No depression and ε = 0.04 (B) Depression strength β = 0.2 and ε = 0.01. (C) Depression strength β = 0.4 and ε = 0.0025. (D) Expected likelihood p[IR > IRT ∗(∞)] the right input IR is stronger than the left IL based in the limit of an infinite number of samples of the dominance times TR and TR for the parameters in (A) (pink); (B) (magenta); and (C) (red). Other parameters are τ = 50 and κ = 0.5. As opposed to the exponential distribution, (42) is peaked away from zero at Tj = kσ, which is also the mean of the distribution. Therefore, two distributions of dominance times with different means will be more easily discerned from one another. We show this in Fig. 8(B) by superimposing the two distributions from Fig. 7 on top of one another. They clearly separate better than in the probability densities of purely noise-driven switching, shown in Fig. 8(A). As the strength of synaptic depression is increased even further, keeping the mean dominance times (cid:104)TR(cid:105) and (cid:104)TL(cid:105) the same, probability densities separate even further (Fig. 8(C)). We summarize how this sep- aration improves the inference of input contrast difference in Fig. 8(D). As the strength β of depression is increased and noise is decreased, an observer's ability to discern which input was stronger is improved. The likelihood assigned to IR being greater than IL is a sigmoidal function of IR whose steepness increases with β. For no noise, the likelihood function is simply a step function H(IR > IL), implying perfect discernment. 3.4. Analyzing switching in a reduced model. We now perform similar analysis on a reduce competitive network model (6) and extend some of the results for the ring model. One of the advantages is that we can construct an energy function [20], 15 which provides us with intuition as to the exponential dependence of mean dominance times on input strengths in the noise-driven case. In particular, we will analyze (6) where the firing rate function is Heaviside (3), starting with the case of no noise uR = −uR + H(IR − qLuL), τ qR = 1 − qR − βuRqR, uL = −uL + H(IL − qRuR) τ qL = 1 − qL − βuLqL. (43a) (43b) First, we note (43) has a stable winner-take-all solution in the jth population (j = R, L) for Ij > 0 and Ik < 1/(1 + β) (k (cid:54)= j). Second, a stable fusion state exists when both IL, IR > 1/(1 + β). Coexistent with the fusion state, there may be rivalrous oscillations, as we found in the spatially extended system (9). To study these, we make a similar fast-slow decomposition of the model (43), assuming τ (cid:29) 1 to find uj's possess the quasi-steady state uR = H(IR − qLuL), uL = H(IL − qRuR). (44) so we expect uj = 0 or 1 almost everywhere. Therefore, we can estimate the domi- nance time of each stimulus using a piecewise equation for the slow subsystem (cid:26) 1 − qj − βqj 1 − qj τ qj = : uj = 1, : uj = 0, j = R, L. (45) Combining the slow subsystem (45) with the quasi-steady state, we can use self- consistency to solve for the dominance times TR and TL of the right and left popula- tions. We simply note that switches will occur through escape mechanism, when the cross-inhibition between populations becomes weak enough such that the suppressed population's (j) input becomes superthreshold, so Ij = qk. Using (45) as we did in the spatial system, we find (cid:34) (cid:34) β − Id +(cid:112)(β − Id)2 − 4(1 − IR)(1 + β)[(1 + β)IL − 1] β + Id +(cid:112)(β + Id)2 − 4(1 − IL)(1 + β)[(1 + β)IR − 1] 2(1 + β)IL − 2 (cid:35) (cid:35) 2(1 + β)IR − 2 TR = τ ln TL = τ ln , , (46) (47) where Id = (1 + β)[IR − IL]. For symmetric stimuli, IL = IR = I, both (46) and (47) reduce to (cid:34) β +(cid:112)β2 − 4(1 − I)(1 + β)[(1 + β)I − 1] (cid:35) , (48) T = τ ln 2(1 + β)I − 2 using which we can solve for the critical input strength I above which only the fusion state exists I = 2 + β 2(1 + β) , (49) in the case of symmetric inputs. We show in Fig. 9 that this asymptotic approxima- tions (46) and (47) of the dominance times match well with the results of numerical simulations. Levelt's propositions are recapitulated well. 16 Fig. 9. Dominance times computed adiabatically in the noise-free competitive network with depression (43). (A) Plot of dominance times T as a function of the strength of a symmetric input IR = IL = I to the competitive system (43) show fast-slow theory (curve) match numerical simulations (dots) very well. (B) Dominance times TL and TR as a function of right input IR keeping IL = 0.6 fixed as computed by theory (curves) in (46) and (47) fits numerically computed (dots) very well. Other parameters are β = 1 and τ = 50. Fig. 10. Noise induced transitions in depression-free two population network (50). (A) Single realization with IR = IL = I = 0.9 for the right uR (red) and left uL (black) population activities. (B) Mean dominance time (cid:104)T(cid:105) as a function of input strength I computed numerically (red dots) and fit to the theoretically derived exponential function (51). (C) Mean dominance times (cid:104)TR(cid:105) and (cid:104)TL(cid:105) as a function of the right input strength IR while IL = 0.95 is fixed. Other parameters are ε = 0.01. Now, we study noise-induced switching in the competitive network. We can sep- arate timescales to study the effects of depression and noise together. To start, we consider the limit of no depression β → 0, so that uR = −uR + H(IR − uL) + ξR, uL = −uL + H(IL − uR) + ξL, (50a) (50b) where ξj are independent white noise processes with variance ε. We show a single realization of the competitive network in Fig. 10(A). Most of the time, the dynamics remains close to one of the winner-take-all attractors where uj = 1 and uk = 0 (j = R, L and k (cid:54)= j). Occasionally, noise causes large deviations where the suppressed population's activity rises above threshold, causing the once dominant population to then be suppressed. We plot the relationship between the strength of the inputs and the mean dominance times in Fig. 10. Notice, in the case of symmetric inputs 17 IL = IR = I, we can fit this relationship to the exponential (cid:104)T(cid:105) ≈ A exp[B(1 − I)]. (51) To understand why this is so, we can study the energy function associated with the system (50). Notice, this network is essentially the classic two neuron flip-flop Hopfield network. As has been shown before [20], for symmetric inputs the energy function for this network can be defined E[uR, uL] = H(I − uR)H(I − uL) − I [H(I − uR) + H(I − uL)] , (52) so we can compute the energy difference between the winner-take-all and fusion states E[1, 0] = −I, E[1, 1] = 1 − 2I, (53) respectively. By taking the difference between these two quantities, we find ∆E = 1− I, which well approximates the exponential dependence of the dominance times T , as shown by the fit in Fig. 10(B). This provides the intuition as to this relationship. In the same way, we can write down the energy function in the case where the inputs are non-symmetric IR (cid:54)= IL, which is [12] E[uR, uL] = H(IL − uR)H(IR − uL) − ILH(IL − uR) − IRH(IR − uL), (54) which means the energy depth of the right and left winner-take-all states are ∆ER = 1 − IL and ∆EL = 1 − IR, respectively. Thus, as we observe in Fig. 10(C), we expect the dominance time of each population to depend upon the strength of the other stimulus according to (cid:104)TR(cid:105) ≈ AR exp [BR(1 − IL)] , (cid:104)TL(cid:105) ≈ AL exp [BL(1 − IR)] . (55) Interestingly, this simple model agrees well with the qualitative predictions of Levelt propositions (i-iv) in this high contrast input regime. Now, we will see how including synaptic depression in the model generates distributions of dominance times that are more similar to those observed experimentally [17, 30, 6]. Finally, we show that the network with depression and noise generates gamma distributed dominance times, as the spatially extended system does. In addition, we provide some analytic intuition as to how gamma distributed dominance times may arise in the fast slow system. First, we display as single realization of the network (6) in Fig. 11(A) along with a plot of an adiabatically computed energy function E[uR, uL, qR, qL] for the system. To compute the energy function, we first note that in the limit of slow depression recovery time τ (cid:29) 1, we can assume the energy of the system will be defined simply by (52) augmented by the synaptic scalings imposed by qR and qL [34]. In the fully general case, where inputs IR and IL may be asymmetric we have E[uR, uL, qR, qL] =H(IL − qRuR)H(IR − qLuL) − IL qR H(IL − qRuR) − IR qL H(IR − qLuL). (56) A similar energy function was previously used in a model with spike frequency adap- tation [37]. Here, we are able to derive the energy function from the model (6). Therefore, the energy gap between a winner-take-all state and the fusion state will 18 Fig. 11. (A) Single realization of the network (6) with depression and noise. Activity variables uR (black) and uL (blue) stay close to attractors at 0 and 1, aside from depression or noise induced switching. Depression variables qR (red) and qL (green) slowly exponentially change in response to the states of uR and uL. (B) Probability density p(T ) of dominance times T sampled over 1000s, well fit by a gamma distribution (42). Parameters are ε = 0.036, β = 0.2, τ = 50, and I = 0.8. Fig. 12. Distribution of dominance times for (A) right and (B) left populations fit with red and blue gamma distributions (42) respectively, in the network (6) with depression and noise in the case of asymmetric inputs IR = 0.82 and IL = 0.78, sampled over 1000s. The right population has longer dominance times. Other parameters are β = 0.2, τ = 50, and ε = 0.036. be time-dependent, varying as the synaptic scaling variables qR and qL change. The energy difference between the right dominant state and fusion is ∆ER(t) = 1 − IL qR(t) , ∆EL(t) = 1 − IR qL(t) , (57) for the right and left population respectively. Notice that dominance times of stochastic switching (Fig. 11) in (6) are dis- tributed roughly according to a gamma distribution (42). Superimposing the prob- ability density of right (left) dominance times on the left (right) probability density, we see they are reasonably separated. Using the analysis we performed for the spa- tially extended system, we could also show that depression improves discernment of the input contrast difference. Mainly here, we wanted to provide a justification as to the relationship between input strength and mean dominance times. Using en- 19 Fig. 13. Perceptual tristability. Examples of images with three possible interpretations. (A) Three overlapping gratings. Redrawn with permission from [38]. (B) 'Mother, father, and daughter.' Redrawn with permission from [16]. Staring at tristable images for long enough leads to the perception switching between the three possible interpretations. (C) Numerical simulation of showing the activity variables u1, u2, u3 and the second synaptic scaling variable q2 (cyan) of the three population network (58) driven by symmetric stimulus I = 0.7. (D) Relationship between the strength of the stimulus I and the dominance times T computed using fast-slow analysis (black) and numerics (red dots). Other parameters are β = 1 and τ = 50. ergy arguments, we have provided reasoning behind why Levelt's propositions are still preserved in this model, when noise is included, even when switches are noise- induced. Increasing one input leads to a reduction in the energy barrier between the other population's winner-take-all state and the fusion state. This leads to the other population's dwell time being shorter. 3.5. Switching between three percepts. Finally, we will compare the trans- fer of information in competitive networks that process more than two inputs. Re- cently, experiments have revealed that perceptual multistability can switch between three or four different percepts [16, 9, 38, 22]. In particular, the work of [38] charac- terized some of the switching statistics during the oscillations of perceptual tristabil- ity. Fig. 13(A,B) shows examples of tristable percepts. Since dominance times are gamma distributed and there is memory evident in the ordering of percepts, the pro- cess is also likely governed by some slow adaptive process in addition to fluctuations. We will pursue the study of perceptual tristability in a competitive neural network 20 model with depression and noise. In the case of three different percepts, a Heaviside firing rate (3), and symmetric inputs I1 = I2 = I3 = I, we study the system u1 = −u1 + H(I − q2u2 − q3u3), u2 = −u2 + H(I − q1u1 − q3u3), u3 = −u3 + H(I − q1u1 − q2u2), τ qj = 1 − qj − βujqj, j = 1, 2, 3. (58a) (58b) (58c) (58d) We are interested in rivalrous oscillations, which do arise in this network for certain parameter regimes (Fig. 13(C)). As per our previous analysis, we perform a fast-slow decomposition of our system. In the case of symmetric inputs, we use our techniques to compute the dominance time T of a population as it depends on input strength I. Our analysis follows along similar lines to that carried out for the two population network, where we assume τ (cid:29) 1. We find (1 − I)(1 + β) +(cid:112)(1 + β)(1 − I)[3I(1 + β) + β − 3] (cid:35) (cid:34) T = τ ln 2[(1 + β)I − 1] , (59) which compares very well with numerically computed dominance times in Fig. 13(D). While perceptual tristability has not been explored very much experimentally [16, 38, 22], observations that have been made suggest that relationships between mean dominance time and input contrast may be similar to the two percept case [22]. In our model, we see that as the input strength is increased, dominance times decrease. One other important point is that percept dominance occurs in the same order every time (Fig. 13(C)): one, two, three. There are no "switchbacks." We will show that this can occur in the noisy regime, which degrades information transfer. Now, we seek to understand how noise alters the switching behavior when added to the deterministic network (58). Thus, we discuss the three population competitive network with noisy depression u1 = −u1 + H(I − q2u2 − q3u3), u2 = −u2 + H(I − q1u1 − q3u3), u3 = −u3 + H(I − q1u1 − q2u2), τ qj = 1 − qj − βujqj + ξj, j = 1, 2, 3, (60a) (60b) (60c) (60d) where ξj are identical independent white noise processes with variance ε. In Fig. 14, we show the noise in (60) degrades two pieces of information carried by dominance switches: the switching time and the direction of switching. Notice that as the ampli- tude of noise ε is increased, the dominance times become more spread out. Thus, there is a less precise characterization of the input strength in the network. Concerning the direction of switching, we see that the introduction of noise makes "switch backs" more likely. We define a "switch back" as a series of three percepts that contains the same percept twice (e.g. 1 → 3 → 1). This as opposed to a "switch forward," which contains all three percepts (e.g. 1 → 3 → 2). Statistics like these were analyzed from psychophysical experiments of perceptual tristability, using an image like Fig. 13(A) [38]. The main finding of [38] concerning this property is that switch forwards occurred more often than chance would suggest. Therefore, they proposed that some slow process may be providing a memory of the previous image. We suggest short term depression as a candidate substrate for this memory. As seen in Fig. 14, the bias 21 Fig. 14. Noise degrades two sources of information provided by dominance switches. (A) In the absence of noise, switches always move "forward," so that the previous percept perfectly predicts the subsequent percept. Dominance times accumulate at a single value too. (B) For slightly higher levels of noise ( ε = 0.0002), "switch backs" can occur where the subsequent percept is the same √ as the previous percept. Also, the distribution of dominance times spreads. (C) For stronger noise ( ε = 0.0004). Other parameters are I0 = 0.6, β = 1, and τ = 50. √ Fig. 15. The probability of a switch being in the forward direction in simulations of (60) as a function of the amplitude ε of noise. As ε increases, network switches behave in more of a Markovian way, not reflecting any memory of the previous percept. Therefore, information of the previous percept is lost as soon as a switch occurs. 22 in favor of switching forward persists even for substantial levels of noise. The idea of short term plasticity as a substrate of working memory was also recently proposed in [35]. Our results extends this idea, suggesting synaptic mechanisms of working memory may be useful in visual perception tasks, such as understanding ambiguous images. In Fig. 15, we show that the process of dominance switching becomes more ε is increased even a modest amount. In the limit Markovian as the level of noise of large noise, the likelihoods of "switch forwards" and "switch backs" are the same. √ 4. Discussion. Mechanisms underlying stochastic switching in perceptual ri- valry have been explored in a variety of psychophysical [17, 30, 6], physiological [31, 5], and theoretical studies [33, 29, 37]. Since psychophysical data is widely accessible, it can be valuable to use the hallmarks of its statistics as benchmarks for theoreti- cal models. For instance, the fact that dominance time distributions are unimodal functions peaked away from zero suggests that some adaptive process must underlie switching in addition to noise [29, 6, 44]. In addition, [36] recently suggested the visual system may sample the posterior distribution of interpretations of bistable im- ages. This type of sampling can be well modeled by attractor networks analogous to those presented here [37]. Therefore, many dominance time statistics from perceptual rivalry experiments can be employed as points of reference for physiologically based models of visual perception. New data now exists concerning tristable images showing this process also likely is guided by a slow adaptive process in addition to fluctuations [38]. We have studied various aspects of competitive neuronal network models of per- ceptual multistability that include short term synaptic depression. First, we were able to analyze the onset of rivalrous oscillations in a ring model with synaptic depression [54, 25]. Stimulating the network with a bimodal input leads to winner-take-all so- lutions, in the form of single bumps, in the absence of synaptic depression. As the strength of synaptic depression is increased, the network undergoes a bifurcation which leads to slow oscillations whose timescale is set by that of synaptic depression. Each stimulus peak is represented in the network by a bump whose dominance time is set by the height of each peak. Thus, synaptic depression reveals information about the stimulus that would otherwise be masked by the lateral inhibitory connectivity of the network. The inclusion of noise in the network leads to dominance times that are exponentially (gamma) distributed in the absence (presence) of synaptic depres- sion. Motivated by recent work exploring how visual perception may exploit Bayesian sampling on posterior distributions [24, 19, 36], we considered the simple task of an observer trying to infer the contrast of stimuli based on dominance times. We found Bayesian sampling of the dominance times discerning input contrast differences better as switches become more depression driven and less noise-driven. Thus, short term depression improves information transfer of networks that process ambiguous images in multiple ways. We also used energy methods in simple space-clamped neural network models to understand how a combination of noise and depression interact to produce switching in competitive neural networks. Using the energy function derived by Hopfield for analog neural networks, we justify the exponential dependence of dominance times upon input strength in purely noise-driven switching. Studying an adiabatically de- rived energy function for the case of slow depression, we also show how depression works to reduce the energy barrier between winner-take-all states, leading to the slow timescale that defines the peak in depression-noise generated switches. Finally, using a three population space-clamped neural network, we analyzed depression and noise 23 generated switching that may underlie perceptual tristability. We found this network also sustained some of the same relationships between input contrast and dominance times as the two population network. Also, we found that when switches are gen- erated by depression there is an ordering to the population dominance that is lost when switches are noise generated. This is due to the memory generated by short term depression [35], so the switching process is non-Markovian due to the inherent slow timescale in the background. However, even small amounts of noise can wash this memory away. Thus, since recent psychophysical experiments reveal a non-Markovian property to percept ordering, this provides further support for the idea that a slow adaptive process underlies percept switching. Note to analytically study the relationship between dominance times and input contrast in the noisy system, we resorted to a simple space-clamped neural network. In future work, we plan to develop energy methods for spatially extended systems like (35). Such methods have seen success in analyzing stochastic partial differential equa- tion models such as Ginzburg-Landau models [14]. Energy functions have recently been developed for neural field models, but have mostly been studied as a means of determining global stability in deterministic systems [53, 28, 40]. We proposed that by deriving the specific potential energy of spatially extended neural fields, it may be possible to approximate the transition rates of solutions from the vicinity of one attractor to another. In the system (35), there should be some separatrix between the two winner-take-all states that must be crossed in order for a transition to occur. The least action principle states that there is even a specific point on this separatrix through which the dynamics most likely flows [14]. Finding this with an energy func- tion in hand would be straightforward would allow us to relate the parameters of the model to the distribution of dominance times. This would provide a better theoretical framework for interpreting data concerning rivalry of spatially extended images, such as those that produce waves [50]. REFERENCES [1] L. F. Abbott, J. A. Varela, K. Sen, and S. B. Nelson, Synaptic depression and cortical gain control., Science, 275 (1997), pp. 220 -- 224. [2] S. Amari, Dynamics of pattern formation in lateral-inhibition type neural fields., Biol Cybern, 27 (1977 Aug 3), pp. 77 -- 87. [3] R. Ben-Yishai, R. L. Bar-Or, and H. Sompolinsky, Theory of orientation tuning in visual cortex, Proc. Natl. Acad. Sci. USA, 92 (1995), pp. 3844 -- 3848. [4] R. Blake, A neural theory of binocular rivalry, Psychol Rev, 96 (1989), pp. 145 -- 67. [5] R. Blake and N. Logothetis, Visual competition, Nature Reviews Neuroscience, 3 (2002), pp. 1 -- 11. [6] J. W. Brascamp, R. van Ee, A. J. Noest, R. H. A. H. Jacobs, and A. V. van den Berg, The time course of binocular rivalry reveals a fundamental role of noise, J Vis, 6 (2006), pp. 1244 -- 56. [7] P. C. Bressloff and J. D. Cowan, An amplitude equation approach to contextual effects in visual cortex, Neural Comput, 14 (2002), pp. 493 -- 525. [8] A. Buckthought, J. Kim, and H. R. Wilson, Hysteresis effects in stereopsis and binocular rivalry, Vision Res, 48 (2008), pp. 819 -- 30. [9] G. Burton, Successor states in a four-state ambiguous figure, Psychonomic Bulletin and Re- view, 9 (2002), pp. 292 -- 297. [10] O. Carter, T. Konkle, Q. Wang, V. Hayward, and C. Moore, Tactile rivalry demonstrated with an ambiguous apparent-motion quartet, Curr Biol, 18 (2008), pp. 1050 -- 4. [11] F. S. Chance, S. B. Nelson, and L. F. Abbott, Synaptic depression and the temporal response characteristics of v1 cells, J Neurosci, 18 (1998), pp. 4785 -- 99. [12] T. Chen and S. I. Amari, Stability of asymmetric hopfield networks, IEEE Trans Neural Netw, 12 (2001), pp. 159 -- 63. 24 [13] D. Deutsch, An auditory illusion, Nature, 251 (1974), pp. 307 -- 309. [14] W. E, W. Ren, and E. Vanden-Eijnden, Minimum action method for the study of rare events, Communications on Pure and Applied Mathematics, 57 (2004), pp. 637 -- 656. [15] D. Ferster and K. D. Miller, Neural mechanisms of orientation selectivity in the visual cortex, Annu Rev Neurosci, 23 (2000), pp. 441 -- 71. [16] G. H. Fisher, "mother, father, and daughter": a three-aspect ambiguous figure, Am J Psychol, 81 (1968), pp. 274 -- 7. [17] R. Fox and J. Herrmann, Stochastic properties of binocular rivalry alternations, Attention, Perception, & Psychophysics, 2 (1967), pp. 432 -- 436. [18] M. Hausser and A. Roth, Estimating the time course of the excitatory synaptic conductance in neocortical pyramidal cells using a novel voltage jump method, J Neurosci, 17 (1997), pp. 7606 -- 25. [19] J. Hohwy, A. Roepstorff, and K. Friston, Predictive coding explains binocular rivalry: an epistemological review, Cognition, 108 (2008), pp. 687 -- 701. [20] J. J. Hopfield, Neurons with graded response have collective computational properties like those of two-state neurons, Proc Natl Acad Sci U S A, 81 (1984), pp. 3088 -- 92. [21] J.-M. Hupe, Dynamics of m´enage `a trois in moving plaid ambiguous perception, J Vis, 10 (2010), p. 1217. [22] J.-M. Hup´e and D. Pressnitzer, The initial phase of auditory and visual scene analysis, Philos Trans R Soc Lond B Biol Sci, 367 (2012), pp. 942 -- 53. [23] Y. Kawaguchi and Y. Kubota, Gabaergic cell subtypes and their synaptic connections in rat frontal cortex, Cereb Cortex, 7 (1997), pp. 476 -- 86. [24] D. Kersten, P. Mamassian, and A. Yuille, Object perception as bayesian inference, Annu Rev Psychol, 55 (2004), pp. 271 -- 304. [25] Z. P. Kilpatrick and P. C. Bressloff, Binocular rivalry in a competitive neural network with synaptic depression, SIAM J Appl Dyn Syst, 9 (2010), pp. 1303 -- 1347. [26] Z. P. Kilpatrick and P. C. Bressloff, Effects of synaptic depression and adaptation on spatiotemporal dynamics of an excitatory neuronal network, Physica D: Nonlinear Phe- nomena, 239 (2010), pp. 547 -- 560. Mathematical Neuroscience. [27] Z. P. Kilpatrick and P. C. Bressloff, Spatially structured oscillations in a two-dimensional excitatory neuronal network with synaptic depression, J Comput Neurosci, 28 (2010), pp. 193 -- 209. [28] S. Kubota and K. Aihara, Analyzing global dynamics of a neural field model, Neural Pro- cessing Letters, 21 (2005), pp. 133 -- 141. [29] C. R. Laing and C. C. Chow, A spiking neuron model for binocular rivalry, J Comput Neurosci, 12 (2002), pp. 39 -- 53. [30] S. R. Lehky, Binocular rivalry is not chaotic, Proceedings of the Royal Society of London. Series B: Biological Sciences, 259 (1995), pp. 71 -- 76. [31] D. A. Leopold and N. K. Logothetis, Activity changes in early visual cortex reflect monkeys' percepts during binocular rivalry, Nature, 379 (1996), pp. 549 -- 53. [32] W. J. M. Levelt, On binocular rivalry, Institute for Perception RVO -- TNO, Soesterberg, The Netherlands, 1965. [33] K. Matsuoka, The dynamic model of binocular rivalry, Biol Cybern, 49 (1984), pp. 201 -- 8. [34] J. F. Mejias, H. J. Kappen, and J. J. Torres, Irregular dynamics in up and down cortical states, PLoS One, 5 (2010), p. e13651. [35] G. Mongillo, O. Barak, and M. Tsodyks, Synaptic theory of working memory, Science, 319 (2008), pp. 1543 -- 6. [36] R. Moreno-Bote, D. C. Knill, and A. Pouget, Bayesian sampling in visual perception, Proc Natl Acad Sci U S A, 108 (2011), pp. 12491 -- 6. [37] R. Moreno-Bote, J. Rinzel, and N. Rubin, Noise-induced alternations in an attractor net- work model of perceptual bistability, J Neurophysiol, 98 (2007), pp. 1125 -- 39. [38] M. Naber, G. Gruenhage, and W. Einhauser, Tri-stable stimuli reveal interactions among subsequent percepts: rivalry is biased by perceptual history, Vision Res, 50 (2010), pp. 818 -- 28. [39] J. Orbach, D. Ehrlich, and H. A. Heath, Reversibility of the necker cube. I. an examination of the concept of "satiation of orientation", Percept Mot Skills, 17 (1963), pp. 439 -- 58. [40] M. R. Owen, C. R. Laing, and S. Coombes, Bumps and rings in a two-dimensional neural field: splitting and rotational instabilities, New Journal of Physics, 9 (2007), p. 378. [41] D. Pressnitzer and J.-M. Hup´e, Temporal dynamics of auditory and visual bistability reveal common principles of perceptual organization, Curr Biol, 16 (2006), pp. 1351 -- 7. [42] J. Seely and C. C. Chow, Role of mutual inhibition in binocular rivalry, J Neurophysiol, 106 (2011), pp. 2136 -- 50. 25 [43] A. Shpiro, R. Curtu, J. Rinzel, and N. Rubin, Dynamical characteristics common to neu- ronal competition models, J Neurophysiol, 97 (2007), pp. 462 -- 73. [44] A. Shpiro, R. Moreno-Bote, N. Rubin, and J. Rinzel, Balance between noise and adaptation in competition models of perceptual bistability, J Comput Neurosci, 27 (2009), pp. 37 -- 54. [45] M. Tsodyks, K. Pawelzik, and H. Markram, Neural networks with dynamic synapses, Neural Comput., 10 (1998), pp. 821 -- 835. [46] M. V. Tsodyks and H. Markram, The neural code between neocortical pyramidal neurons depends on neurotransmitter release probability, Proc. Natl. Acad. Sci. USA, 94 (1997), pp. 719 -- 723. [47] R. van Ee, Stochastic variations in sensory awareness are driven by noisy neuronal adaptation: evidence from serial correlations in perceptual bistability, J Opt Soc Am A Opt Image Sci Vis, 26 (2009), pp. 2612 -- 22. [48] X. Wang and J. Rinzel, Alternating and synchronous rhythms in reciprocally inhibitory model neurons, Neural Computation, 4 (1992), pp. 84 -- 97. [49] H. R. Wilson, Computational evidence for a rivalry hierarchy in vision, Proc Natl Acad Sci U S A, 100 (2003), pp. 14499 -- 503. [50] H. R. Wilson, R. Blake, and S. H. Lee, Dynamics of travelling waves in visual perception, Nature, 412 (2001), pp. 907 -- 10. [51] H. R. Wilson and J. D. Cowan, A mathematical theory of the functional dynamics of cortical and thalamic nervous tissue, Biol Cybern, 13 (1973), pp. 55 -- 80. [52] J. M. Wolfe, Stereopsis and binocular rivalry, Psychol Rev, 93 (1986), pp. 269 -- 82. [53] S. Wu, S.-I. Amari, and H. Nakahara, Population coding and decoding in a neural field: a computational study., Neural Comput, 14 (2002), pp. 999 -- 1026. [54] L. C. York and M. C. W. van Rossum, Recurrent networks with short term synaptic depres- sion, J Comput Neurosci, 27 (2009), pp. 607 -- 20. [55] W. Zhou and D. Chen, Binaral rivalry between the nostrils and in the cortex, Curr Biol, 19 (2009), pp. 1561 -- 5. 26
1703.04176
1
1703
2017-03-12T21:04:39
Locally embedded presages of global network bursts
[ "q-bio.NC", "math.DS", "nlin.CD" ]
Spontaneous, synchronous bursting of neural population is a widely observed phenomenon in nervous networks, which is considered important for functions and dysfunctions of the brain. However, how the global synchrony across a large number of neurons emerges from an initially non-bursting network state is not fully understood. In this study, we develop a new state-space reconstruction method combined with high-resolution recordings of cultured neurons. This method extracts deterministic signatures of upcoming global bursts in "local" dynamics of individual neurons during non-bursting periods. We find that local information within a single-cell time series can compare with or even outperform the global mean field activity for predicting future global bursts. Moreover, the inter-cell variability in the burst predictability is found to reflect the network structure realized in the non-bursting periods. These findings demonstrate the deterministic mechanisms underlying the locally concentrated early-warnings of the global state transition in self-organized networks.
q-bio.NC
q-bio
Locally embedded presages of global network bursts Satohiro Tajima1,2,3,*, Takeshi Mita4, Douglas J. Bakkum5, Hirokazu Takahashi4, and Taro Toyoizumi3 1) Department of Neuroscience, University of Geneva, CMU, Rue Michel-Servet 1, Genève, 1211, Switzerland. 2) JST PRESTO, Japan Science and Technology Agency, 4-1-8 Honcho, Kawaguchi, Saitama, 332-0012, Japan. 3) RIKEN Brain Science Institute, 2-1, Hirosawa, Wako, Saitama, 351-0198, Japan. 4) Research Center for Advanced Science and Technology, the University of Tokyo, 4-6-1, Komaba, Meguro-ku, Tokyo, 153-8904, Japan. 5) ETH Zurich, Department of Biosystems Science and Engineering, Basel, Switzerland. *[email protected], CMU, Rue Michel-Servet 1, Genève, 1211, Switzerland. Abstract Spontaneous, synchronous bursting of neural population is a widely observed phenomenon in nervous networks, which is considered important for functions and dysfunctions of the brain. However, how the global synchrony across a large number of neurons emerges from an initially non-bursting network state is not fully understood. In this study, we develop a new state-space reconstruction method combined with high-resolution recordings of cultured neurons. This method extracts deterministic signatures of upcoming global bursts in "local" dynamics of individual neurons during non-bursting periods. We find that local information within a single-cell time series can compare with or even outperform the global mean field activity for predicting future global bursts. Moreover, the inter-cell variability in the burst predictability is found to reflect the network structure realized in the non-bursting periods. These findings demonstrate the deterministic mechanisms underlying the locally concentrated early-warnings of the global state transition in self-organized networks. 1 Introduction The collective firing dynamics of neural population have been related to emergent functions and dysfunctions in the brain. Unraveling the structure of dynamics emerging from the collective behavior in neural networks is an ultimate issue in studies of complex systems (1). The simplest and most profound example of collective dynamics is synchronous burst: the simultaneous occurrence of closely-spaced action potentials ("bursts") across a large number of neurons in the network. Within neural circuits in vivo, the synchronous bursts are believed to serve important roles in information storage and transmission (2, 3) as well as in disease states including epileptic seizures (4–6). Intriguingly, neurons cultured in vitro also display spontaneous synchronous bursts akin to those observed in vivo (7–9), indicating that the collective bursting is a universal behavior of neural networks in a wide variety of preparations. Classic biophysical models of the synchronized bursts assume broadly diverging synaptic interaction that propagates activity of a single cell to the remaining neural population (4, 10–12). This view has been partially supported by experiments demonstrating that stimulating a single neuron can change the large-scale dynamics of neural populations (13– 15). These models and experiments suggest a close link between single neuron activity and the global state of a neural population. On the other hand, the previous stochastic models do not specify when a global burst occurs; forecasting a spontaneous occurrence of synchronous burst from the preceding individual neural activities in non-bursting states remains a highly challenging issue, particularly in real nervous systems which feature the heterogeneity of local-global coupling (e.g., non-uniform coupling strength between individual neural activity and the overall firing of the population) (16). Indeed, recent studies demonstrate potentially diverse dynamical mechanisms accompanied by complex but reproducible patterns of activity sequences leading to the synchronous bursts (17, 18), implying the existence of nonlinear deterministic mechanisms around the burst initiations. The present paper investigates the deterministic aspects of burst initiation. In particular, we question whether individual neuron dynamic in non-bursting periods can predict an upcoming synchronous burst in cultured neural population. To cope with the heterogeneous nature of local-global coupling, we introduce a model-free method to quantify the deterministic relationships among individual neuron traces and the global state, based on a nonlinear state-space reconstruction technique. Using this method, we report that dynamic of only one neuron can robustly predict the upcoming synchronous burst in the neural population, at high signal-to-noise ratio. Surprisingly, the burst predictability with an appropriately-chosen neuron even outperforms that with the mean field activity of neural population, demonstrating that macroscopic fluctuation of neural population is better predicted by the microscopic dynamic of a specific single cell, rather than the macroscopic state itself. This apparently counterintuitive property is explained based on mathematical property termed 'embedding' in nonlinear dynamical systems and the heterogeneous input-output interactions among neurons in the network shaping non- bursting spontaneous dynamics. Results Spontaneous synchronous bursts in cultured neurons Rat cortical neurons were grown on high-dense CMOS-based multi-electrode arrays for 12-41 days-in-vitro (MEAs; Fig. 1a) (19, 20). The devices allow us to simultaneously stimulate and record from multiple neurons lying on the array surface at a high spatiotemporal resolution (Methods). Each recording yielded simultaneous traces of 60–98 cells at a temporal resolution of 2 kHz. Figure 1b shows the cell distribution recorded with a representative array preparation. In our preparations of sparse culturing, the cell bodies were well isolated, and the action potential of each cell was identified accurately based on its spatial location and waveform. Spike action-potentials were detected using standard LimAda method (21). We first analyzed the spontaneous activities of the recorded neural population. During the spontaneous firing, the neurons showed intermittent synchronous bursts, which were interleaved with silent periods (Fig. 1c). To describe the macroscopic network dynamics, we defined population activity by the mean firing rate across all recorded neurons (Fig. 1d). The synchronous bursting period were defined by time bins in which the population activity exceeded a threshold determined for each preparation, according to the previously established method based on inter-spike interval distributions (20) (Fig. 1d, inset). The timing of synchronous burst occurrence did not show any clear periodicity. Duration of bursts varied from about 200 to 800 ms. The different preparations showed distinct complex patterns of collective activity dynamics when visualized with standard dimensionality reduction techniques such as principal component analysis (Supplementary Fig. S1). 2 Predicting bursts To capture the potentially complex temporal structures in neural activities around bursting dynamics, we developed a model-free method that is able to quantify the general relationships between the network and single-element dynamics in high-dimensional nonlinear systems. Because the method features tracing the reconstructed state-space trajectories in the retrograde direction, from now on we term the method as trace-back embedding (TBE). The TBE extends the nonlinear forecasting techniques based on state-space reconstruction using the delay-embedding theorems in deterministic dynamical systems (22, 23). Compared to the previous nonlinear forecasting methods (24–26), it has two additional advantages. First, the method focuses particularly on the relationships between a global mean-field variable and local variables. Second, the method quantifies the predictability of a global synchronous burst based on temporally distant dynamics of each local variable. The basic protocol of TBE is summarized below. Figure 2 illustrates the TBE protocol (further details are provided in Methods). We first reconstruct the dynamics of population activity and each single-neuron activity respectively in delay-coordinate state-spaces (Fig. 2a, b): bt d = (bt, bt-τ …, bt-(d-1)τ), xi,t d = (xi,t, xi,t-τ …, xi,t-(d-1)τ), (1) (2) d and xi,t where bt, and xi,t are the population mean-field and neuron i's activity at time t, respectively; d denotes the embedding dimensions (number of delay-coordinates); τ is the unit delay size. In this study, the local time series xi,t was defined by each neuron i's spike train smoothed with a Gaussian kernel, and bt was defined by the mean of xi,t over all the d as the "global" and "local" state simultaneously recorded neurons (Methods). For convenience, we refer to bt trajectories, respectively. The TBE method measures the accuracy of synchronous-burst prediction by seeking "presages" of global bursts in each local state trajectories, based on the accuracy of forecasting using a nearest-neighbor model (Fig. 2b, see also the figure legend for a step-by-step description of the protocol). The rationale behind the TBE protocol is as follows: if neuron i has enough information to predict the future synchronous burst occurring after a time span –Δt (–Δt>0), neuron i's state up to time –Δt before the targeted burst should be already differentiated from the states without any future bursting, forming a cluster of states that deviate from those predicting non-bursting periods in the reconstructed state space. Similarly, we define the accuracy of "postdiction" of bursts (i.e., detecting a burst event based on neural activities after its occurrence) with the same protocol as that for prediction except for using time span Δt with a positive value. We iterated this procedure for all the neurons to characterize the burst predictabilities in individual neurons. In addition, we quantified the "self-predictability" of the mean-field activity by replicating the above procedure based on its own (global) state trajectory, bt, instead of a local state trajectory in individual neurons, xi,t. Using this method, we investigated the predictability of synchronous bursts based on the individual neural activities or the population mean-field. In particular, we addressed two key questions: (i) how accurately can each single neuron predict the future synchronous bursting events, and (ii) if a subset of neurons predicts synchronous bursts better than others, how are they related to the underlying neural network structure? Single-neuron dynamics can predict global network bursts First, we asked how accurately individual neurons predict (or postdict) the occurrence of synchronous bursts, and how their predictions compare to that based on the population mean-field, which defines synchronous burst. We quantified the rate of successful synchronous-burst predictions in each single neuron with keeping the false-positive rate at 5% (Methods). Interestingly, in some cells (e.g., Fig. 3a), the rates of successful burst predictions with the TBE method could be relatively high (>50% of trials) even when the network appeared to be silent in terms of the global mean-field firing rate (Fig. 3d, – 1000 ~ –500ms). Corresponding to this, the success rate of TBE-based burst detection using a single neuron dynamics could be even higher than that with the global mean field dynamics (Fig. 3b). The predictability of burst occurrence varied among neurons, suggesting the heterogeneity of burst predictability across neurons. However, the accuracy of predictions in individual neurons were highly consistent among different sets of burst trials within each preparation (ρ=0.79~0.93, P<2×10-23, Spearman's rank correlation between the first and the second halves of trials; Fig. 3e), demonstrating that this cell-to-cell variability reflects a robust property of each neuron. Across all the preparations, about 1/3 of neurons were found to predict synchronous bursts robustly better than population mean-field over some prediction span, Δt (Fig. 3g). The fact that only a single neuron can compare or even outperform the population mean at predicting the mean-field dynamics may appear surprising, considering that the single-cell activities fluctuate due to spiking much more than the mean-field activity (e.g., compare the black traces in Figs. 3c and 3d). However, it is explained by the delay-embedding theorems (22, 23) that the global state information could be reconstructed from observation of time series in a single variable participating in the dynamical system (Supplementary Fig. S2). The present result suggests that such a mathematical property can be demonstrated in a biological system. Importantly, to reconstruct the global state requires, we need to observe multiple time points rather than the momentary state in the time-series of the observed variable (which is 3 why the theorem is called "delay-embedding" theorem). It suggests that the information about the global state is carried by the temporal patterns including multiple time points, rather than momentary snapshots of activity in neurons. To explore the importance of temporal patterns of activities in predicting bursts compared to the momentary firing rates, we compared the current TBE-based prediction accuracy with an analogous index based on momentary firing rates, xi,t or bt (the red traces in Figs. 3c and 3d, respectively; see also Methods). In single neurons, the TBE method tended to provide more accurate prediction than the momentary-activity-based method (P<5×10-14 in all the individual preparations, sign test with paired samples; Fig. 3f). This superiority of TBE to the momentary firing rate is not likely to be due to the difference in the sampled spike numbers because it remained even if we lengthened the bin size for computing activity so as to have the same total duration as used in TBE (P<6×10-3 in all the preparations except for Chip 2427 [which had much lower firing rate than others], sign test with paired samples; Fig. 3f, the inset). Finally, when we used the momentary firing rates instead of the TBE with temporal sequence, the fraction of cells whose burst detectability outperformed that of the mean field reduced from ~1/3 to ~1/5 (Fig. 3g). Together, these results suggest that the temporal patterns in neural activity, not only momentary activity, are crucial for early detection of synchronous bursts. Burst predictability reflects network structures realized in non-bursting periods Next, we examined whether the predictability of synchronous bursts corresponds to any specific network structure defined by synaptic interactions. The structural correspondences of burst predictability were investigated in three steps (Figs. 4): we first re-analyzed the non-bursting spontaneous activities to infer the directed causal network structure; next, the inferred causal networks were verified by comparing them to synaptic interactions measured by an additional electrical stimulation experiment with the same preparations; finally, we related each neuron's network connectivity with its burst prediction accuracy. In the causal network analysis, we applied a previously proposed method that is capable of detecting weak nonlinear coupling (26) (Methods). Theoretically, the method detects causal interactions including indirect ones (27). Applying this method to all the neuron pairs showing non-bursting spontaneous activities yielded a matrix for each preparation that represents pairwise, directed interactions within each preparation (Fig. 4a). The relevance of the inferred causal couplings to synaptic interactions among neurons was verified using independent data in which synaptic interactions were directly measured (Fig. 4a, b). The present CMOS-based MEA system allowed us to stimulate the neural tissue local to given electrodes at sub-millisecond accuracy, directly activating a single neuron's soma at a high temporal precision with few artifact (19). A subset of neurons was electrically stimulated while the evoked activities in other neurons were monitored simultaneously. According to the evoked response latencies, we identified the short-latency (< 10 ms) and long-latency (≥ 10 ms) downstream cells for each stimulated neuron, between which the latter was expected to include more multi-synaptic interactions and thus weaker causal effects. As expected, we confirmed that the causality inferred based on non-bursting spontaneous data consistently reflected the identified short- and long-latency interactions (Fig. 4b). Specifically, the inferred causality differentiated the short- and long-latency downstream cells from the remaining population (P = 0.0001, sign test, data pooled across the all preparations). Based on these confirmations, we used the inferred causality as proxies for the effective synaptic interactions in the network during the non-bursting periods. Finally, we compared these causal network structures and the burst predictabilities, finding that the burst predictability in each neuron displayed a positive correlation to the average inferred causality between it and the other neurons (Figs. 4d and 4e). We classified putative excitatory and inhibitory neurons based on the spike waveforms (Methods). Interestingly, the excitatory neurons predicted bursts better than the inhibitory neurons when we compared the neurons showing strong causality (P<0.002, both for the causality for input and output connections, rank sum test), but such cell-type dependency was unclear in the neurons having weak interactions (P>0.5; Figs. 4d). These results suggest that the stronger excitatory network interaction with other cells underlie the higher burst predictability. The above results together led us to conclude that the cell-to-cell variability in burst predictability is likely to reflect the heterogeneous network structures that shape non-bursting activity dynamics. Discussion In this study, we have focused on the relationship between local and global neural properties by analyzing how single cell activity predicts the population mean-field dynamics. The newly introduced TBE method revealed the heterogeneity of neurons also in the predictability of upcoming occurrence of a synchronous burst of a neural population, which is temporally distant in order of hundred milliseconds. The bursts were predicted accurately by dynamics of some single neurons. Remarkably, such neuron even outperformed the population mean-field dynamics itself in terms of prediction accuracy, particularly when we extract the information from the temporal patterns of neural activity rather than the momentary firing rates. From causal network analysis combined with the electrical stimulation experiment, the 4 heterogeneity of burst predictability was explained by the structures of neural networks: the "hub"-like neurons having stronger synaptic interactions with other cells can predict the upcoming global network burst better than others. In addition, the present results based on the cultured neurons indicate the heterogeneous relationships between single neurons and population activity are basic properties observed in the nervous network that is self-organized without external stimulus. Classic models of the synchronized population burst assume diverging interaction that propagates activity of a cell to the other cells (4, 10–12). A variant of this model features "avalanche" like bursting sequences, which models synchronous burst with noise being amplified and propagated through the network (9, 18, 28–32). These models, focusing mainly on the stochastic aspects of burst initiation and propagation, have made successes in explaining a variety of statistical properties of bursting neural networks. On the other hand, it has been challenging to predict the occurrence of synchronous burst, partially because they can be triggered by complex and heterogeneous mechanisms (33). The present study, focusing on deterministic aspects of the neural population dynamics that transit from silent state to synchronous burst, demonstrates that the burst occurrence is predictable to the extent based on preceding activity traces in single neurons. This is in line with studies suggesting that the spatiotemporal pattern of spontaneous fluctuation in neural population activity is not purely random but restricted to a limited number of configurations (34, 35). The existence of specific neurons predicting the upcoming burst is also consistent with the previous idea of 'leader neurons' that fire at the beginning of bursts (8, 36). Notably, however, the predictor of the global burst is not identical to the increased activities in those single neurons as expected from previous studies (13, 30, 36), because the magnitude of FR alone is not sufficient to explain the highly accurate predictability based on TBE method (Fig. 3e). Rather, the earliest presages are likely to be embedded in the dynamical patterns of those neurons. On the other hand, it should be noted that not everything is predictable in a deterministic manner before the burst occurrence. A most relevant example is the trial-to-trial fluctuation in burst size (the height of peak population rate during a burst). In contrast to the case of the burst detection, the cell-to-cell heterogeneity in the burst-size predictability (Methods) was not always consistent (ρ = –0.289 ~ –0.077, P>0.4, Spearman's rank correlation between the odd and even trials). Neither the global state could predict the size of upcoming burst (ρ = 0.091 ~ 0.35, P>0.1, Spearman's rank correlation between actual and predicted burst sizes). These facts suggest that the burst size is determined in a different mechanism from that triggering the burst: the former is more likely to follow a stochastic mechanism during the burst (as previously described in probabilistic activity-propagation models (18, 32)) whereas, somewhat counterintuitively, the latter seems to follow more deterministic process. Finally, the predictability of upcoming burst occurrence based on single-cell dynamics suggests an effective methodology to capture the early warnings of population state transition in a variety of networks, including epileptic seizures in the brain (5, 37), rumor propagation in social networks (38–40), or epidemic spreading in human networks (41, 42). In clinical purposes, for example, accurate prediction and characterization of pathologically synchronized neural firing is an important issue (5, 37). Although nonlinear features of field potential sequence have been proposed as useful markers of epileptic state (37, 43–46), the predictability of global dynamics based on single neuron activity, to the best knowledge of ours, has been not thoroughly studied. In particular, the present study suggests that single neuron could even outperform the direct observation of global state itself in predicting the global state dynamics. The success of the present state-space reconstruction-based method implies a new approach to detect the early warnings of transitions in global network state based on local rather than global features. Methods Cell culturing All experimental protocol was approved by the Committee on the Ethics of Animal Experiments at the Research Center for Advanced Science and Technology, the University of Tokyo (Permit Number: RAC130106). A protocol for cell culturing had been reported previously (19). Briefly, E18 Wistar rat cortices were dissociated using trypsin and mechanical trituration. 20-40k/µL neurons and glia were seeded over an area of ~12 mm2 on top of the CMOS chip. Layers of poly (ethyleneimine) followed by laminin were used to adhere cells. Plating media consisted of Neurobasal- B27 supplemented with 10% horse serum and 0.5 mM GlutaMAX during the first 24 hours. Growth media consisted of DMEM supplemented with 10% horse serum, 0.5 mM GlutaMAX, and 1 mM sodium pyruvate. Experiments were conducted inside an incubator to control of environmental conditions (36°C and 5% CO2). CMOS-based recording and stimulation of network activity Cultured neuron activities were simultaneously recorded with high-density microelectrode array as described before (19, 20). Cortical networks were grown over 11,011-electrode CMOS-based MEAs (47, 48), which provide enough spatial and 5 temporal resolution to detect action potentials from any neuron lying on the array: 1.8 × 2.0 mm2 area containing 8.2 × 5.8 μm2 electrodes with 17.8 μm pitch, sampled at 20 kHz. Subsets of 126 electrodes can be read-out (and stimulated) at one time, and electrode selection can be re-configured within a few milliseconds. To identify the locations of neurons growing over the array, a sequence of about one hundred recording configurations were scanned across the whole array while recording spontaneous activity. Locations of active somata were identified because action potentials could be usually detected from multiple nearby electrodes. Electrode selection was then re-configured such that a single electrode was assigned to each identified soma, and spontaneous activities were measured simultaneously from all of these cells. The putative neuron types (excitatory or inhibitory) were estimated based on spike shapes (Supplementary Fig. S3). We could recode from 93, 47, 98, 92 and 53 neurons in Chip 1437, 1440, 1444, 2427 and 2440, respectively. Microstimulation-elicited network activities were then investigated to characterize a pairwise synaptic strength between neurons. An adequate stimulating electrode was explored such that a single target neuron was directly activated through axonal stimulation and that the directly evoked spikes were exclusively measured from the target cell. The directly evoked spikes could be easily distinguished from post-synaptic activations because they were very reliable (i.e., 100 spikes elicited out of 100 stimulation trials) and exhibited a small temporal jitter (Supplementary Fig. S4). The microstimulation was a single, positive-negative, biphasic pulse with a charge-balanced amplitude of ± 300-900 mV and a duration of 200 µs/phase, and was delivered 100 times every 3 sec. Burst detection Burst in spontaneous activity was detected modifying a protocol that was previously established by authors' group (20). The previous study proposes to determine the threshold for burst detection based on the distribution of inter-spike interval (ISI) of consecutive spikes (20). In the bursting neurons, the ISI distribution typically has a bimodal structure whose valley can be used as an objective criterion for burst detection. In this study, we first computed sequences of population firing rate that were normalized such that it ranges from 0 to 1. We defined this as the sequence of global state, bt, where t is the time. This was to use a common burst detection threshold across different preparations of neuron cultures, which generally vary in terms of the absolute firing rate. We used 5 ms bin for the firing rate computation. We next derived the distributions of inverse firing rate over the bins. This yielded an ISI distributions for each preparation, in which we confirmed that all of them had bimodal shapes (Fig. 1d, inset). For each preparation, we selected the smallest ISI providing the valley of distribution as the burst detection threshold. The burst periods were determined by the at least three consecutive sequences of bins that have average ISIs under this threshold, identifying 26 bursts on average for each preparation. We defined the peak burst timing by selecting the center of the bin having the smallest average ISI (i.e., the largest global state) within each burst period. Estimation of synaptic connectivity from electrical stimulation Synaptic connectivity was estimated based on the evoked responses during electrical stimulation experiment. To reliably stimulate a single neuron, we searched a stimulation site around a target neuron in each MEA that could elicit an action potential exclusively at the target neuron. After the careful selection of stimulation sites, we could evoke the action potentials at almost 100% probability for each single stimulation, with very little jitters in the timings of action potentials in the stimulated neurons. We first aligned the spike raster to the timings of stimulation. We next computed the sequence of firing rate xi,t in a way described above. The firing rates before the stimulation (-500 ms < t < -100 ms) were used to produce the reference probability distribution of each neuron's firing rate, P(xi). Next, we computed the probability, P(xi > xi,t), for each time bin centered at t in a short post-stimulation period (2.5 ms < t < 15 ms); the responses during 0 ms < t < 2.5 ms were omitted to eliminate the artifacts of electrical stimulation. We defined the smallest time t that satisfies P(xi > xi,t) < 0.01 in the post-stimulation period as the latency of neuron i. The neurons that had at least one time-bin satisfying this condition were defined as "downstream cells of the stimulated neuron; the neurons that do not have any such time bin were defined as "non-downstream" cells. According to the evoked response latencies, we identified the short- (< 10 ms) and long-latency (≥ 10 ms) downstream cells for each stimulated neuron. The multi-synaptic downstream cells for a stimulated neuron comprised relatively small fraction n (17% on average) of the entire population. This indicates that they are subsets of neurons receiving effectively strong input from the stimulated cell via relatively a small number of path length, although the interaction with longer path- length, which was not detected here, could include the larger fraction of the cell population. Causal network analysis Pairwise causal interaction among neurons was estimated based on a variant of convergent cross-mapping (CCM). CCM was developed by Sugihara et al. (26) as an extension of nonlinear forecasting method of time sequence based on nearest- neighbor models (24, 25, 49). The method is capable of detecting relatively weak causal interactions in deterministic dynamical systems (which can include some stochastic components) if the system and observations are deterministic, the 6 embedding dimension is sufficiently large, and data size is sufficiently large for the given embedding dimension. A variant of CCM that can be used for systems, in which the variables have heterogeneous time-scales (such as in neural system), was developed by the authors (27). Suppose that we want to know the interaction from neuron x to another neuron y. We first reconstruct the state-space 𝑑𝑚𝑎𝑥 = (𝑥𝑡, 𝑥𝑡−𝜏, … , 𝑥𝑡−(𝑑𝑚𝑎𝑥−1)𝜏), where 𝑑𝑚𝑎𝑥 represents the trajectories of each neuron 𝑥 in the delay-coordinates, 𝒙𝑡 maximum number of dimensions (number of delay-coordinates) to be considered, 𝑡 is the time point, and 𝜏 is the unit delay length. We used 𝜏 = 5 ms dmax = 8 in this study. In the embedding-based analyses, we convolved the firing-rate sequence of each neuron with Gaussian kernel that has half-width-of-half-height of 25 ms, and normalized such that each neuron trace ranges from 0 to 1. To avoid a known vulnerability of the state-space reconstruction to the time-scale heterogeneity 𝑑𝑚𝑎𝑥 to a randomized coordinate space by multiplying a square random matrix, 𝑹, from (50), we projected delay vector 𝒙𝑡 𝑑 is constructed by selecting the left to obtain a transformed vector: 𝒙𝑡 𝑑𝑚𝑎𝑥. The causal interaction form neuron x to another neuron y is quantified based the first 𝑑 (≤ 𝑑𝑚𝑎𝑥) components of 𝒙𝑡 on the correlation coefficient, 𝜌𝑦𝑡,𝑦(𝒙𝑡 𝑑), between the true (𝑦𝑡) and forecast (𝑦𝑡(𝒙𝑡 𝑑𝑚𝑎𝑥. A 𝑑-dimensional delay vector 𝒙𝑡 𝑑)) signals, where 𝑑𝑚𝑎𝑥 = 𝑹 𝒙𝑡 𝑦𝑡(𝒙𝑡 𝑑) = ∑ 𝑤(𝒙𝑡′ 𝑑 − 𝒙𝑡 𝑑) 𝑦𝑡′ 𝑡′ s.t. 𝒙𝑡′ 𝑑) 𝑑 ∈𝐵(𝒙𝑡 with the k-nearest neighbor set 𝐵(𝒙𝑡 𝑑) of 𝒙𝑡 𝑑 in the delay-coordinate space. We set 𝑘=4, weight, 𝑤(𝒙𝑡′ 𝑑 − 𝒙𝑡 𝑑) = exp(−𝒙𝑡′ 𝑑 − 𝒙𝑡 𝑑) 𝑑 − 𝒙𝑡 𝑑) exp(−𝒙𝑡′ , 𝑑) 𝛴𝑡′ s.t. 𝒙 𝑑 ∈𝐵(𝒙𝑡 𝑡′ 𝑑 𝑑 and 𝒙𝑡 in the data analysis. Note that 𝜌𝑦𝑡,𝑦(𝒙𝑡 𝑑 − 𝒙𝑡 𝑑 as the square distance between 𝒙𝑡′ and 𝒙𝑡′ prediction. We selected the optimal dimension d=d* so as to maximize the prediction performance, 𝜌𝑦𝑡,𝑦(𝒙𝑡 of d* was distributed from 2 to 4 in the current data. 𝑑) = 1 means perfect 𝑑). Typical value Trace-back embedding method and burst predictability The trace-back embedding (TBE) method is an extension of the above CCM algorithm. The method extends the original CCM (i) by relating the global state to local components while the original protocol only quantified the relationships between local variables, and (ii) by quantifying on the predictability of temporally distant global state based on local variable dynamics. This is done by finding a mapping between the system's global state bt and each neuron 𝑥𝑡−𝛥𝑡 including temporal lag 𝛥t, instead of simultaneous prediction among individual neurons. For burst detection, we set t as the time corresponding to the peak of global state during a burst period, 𝑡 ∈ {𝑡1, … , 𝑡𝑛}, where n is the number of burst trials in the experiment. We first reconstruct the state-space trajectory for neuron x and the normalized global state b in the d = (bt, bt-τ …, bt-(d-1)τ). It is mathematically shown that both (randomized) delay-coordinates, xt of these trajectories reproduce the topological structure (i.e., preserving continuity but allowing deformations) of attractor dynamics that they participate, with which a (near-) future state of each variable is accurately predicted from the current state of other variables, if the dynamics (approximately) follows deterministic mechanisms (22, 23, 26, 27). Practically, however, the accuracy of prediction is limited by various factors including data size, process/observation noises, as well as the uncertainty due to asymmetric causal interactions. d = (xt, xt-τ …, xt-(d-1)τ) and bt The forecast of the global state 𝑏 𝑡 based on a single neuron x with time lag 𝛥𝑡 is given by 𝑏 𝑡 (𝒙𝑡−𝛥𝑡 𝑑 ) = ∑ 𝑑 𝑤(𝒙𝑡′−𝛥𝑡 𝑑 − 𝒙𝑡−𝛥𝑡 ) 𝑏𝑡′ 𝑑 𝑡′ s.t. 𝒙 𝑡′−𝛥𝑡 𝑑 ∈𝐵(𝒙𝑡−𝛥𝑡 ) with the k-nearest neighbor set 𝐵(𝒙𝑡−𝛥𝑡 same weight function, 𝑑 ) of 𝒙𝑡−𝛥𝑡 𝑑 , in the same manner s in CCM. In the data analysis we used k=4 and the 𝑑 𝑤(𝒙𝑡′−𝛥𝑡 𝑑 − 𝒙𝑡−𝛥𝑡 ) = exp(−𝒙𝑡′−𝛥𝑡 𝑑 𝑑 − 𝒙𝑡−𝛥𝑡 ) exp(−𝒙𝑡′−𝛥𝑡 𝑑 𝛴𝒙 𝑑 𝑡′−𝛥𝑡 𝑑 ∈𝐵(𝒙𝑡−𝛥𝑡 ) − 𝒙𝑡−𝛥𝑡 𝑑 . ) The success rate in burst detection was defined by the ratio between the number of burst trials such that 𝑏 𝑡𝑛 > 𝜃𝑥, where 𝜃𝑥 is the threshold of burst detection for neuron x. In the present study, we defined 𝜃𝑥 by the 95 percentile in the null- hypothesis distribution P(𝑏 )), where 𝑡Null is randomly selected time outside the bursting period. The burst- size predictability in each cell was defined by correlation between actual and predicted burst sizes, average over range of time span -200 ms < Δt < -50 ms. 𝑡Null (𝒙𝑡Null−𝛥𝑡 𝑑 7 Similarly, the forecast based on the history of the global state itself is given by the same protocol except for using b instead of x: 𝑏 𝑡 (𝒃𝑡−𝛥𝑡 𝑑 ) = ∑ 𝑑 𝑤(𝒃𝑡′−𝛥𝑡 𝑑 − 𝒃𝑡−𝛥𝑡 ) 𝑏𝑡′ . 𝑑 𝑡′ s.t. 𝒃 𝑡′−𝛥𝑡 𝑑 ∈𝐵(𝒃𝑡−𝛥𝑡 ) This quantifies the self-predictability of burst based on the global state itself, through the success rate compute based on the corresponding threshold 𝜃𝑏, which in the present study was defined by 95 percentile of null-hypothesis distribution P(𝑏 )). 𝑡Null (𝒃𝑡Null−𝛥𝑡 𝑑 Acknowledgements The authors are grateful to Dr. Urs Frey at RIKEN Quantitative Biology Center, Kobe, Japan, and Prof. Andreas Hierlemann at ETH Zurich, Basel, Switzerland, for providing CMOS-based MEA with their technical supports. This work was supported by JST PRESTO (ST), Asahi Glass Foundation, and KAKENHI (26630089), Brain/MINDS from AMED (TT), and RIKEN Brain Science Institute (TT). Conflict of interests The authors declare no competing financial interests. Author contributions Designed research: ST, TM, HT, TT Performed research: - Contributed unpublished reagents/analytic tools: TM, DJB, HT Analyzed data: ST, TT Wrote the paper: ST, TM, HT, TT References 1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. Chialvo DR (2010) Emergent complex neural dynamics. Nat Phys 6(10):744–750. Lisman JE (1997) Bursts as a unit of neural information: Making unreliable synapses reliable. Trends Neurosci 20(1):38–43. Izhikevich EM, Desai NS, Walcott EC, Hoppensteadt FC (2003) Bursts as a unit of neural information: Selective communication via resonance. Trends Neurosci 26(3):161–167. Traub RD, Wong RK (1982) Cellular mechanism of neuronal synchronization in epilepsy. Science (80- ) 216:745– 747. McCormick D, Contreras D (2001) On the cellular and network bases of epileptic seizures. Annu Rev Physiol 63:815–846. Staley KJ, Dudek FE (2006) Interictal Spikes and Epileptogenesis. Epilepsy Curr 6(6):199–202. Feinerman O, Segal M, Moses E (2007) Identification and dynamics of spontaneous burst initiation zones in unidimensional neuronal cultures. J Neurophysiol 97(4):2937–2948. Eckmann JP, Jacobi S, Marom S, Moses E, Zbinden C (2008) Leader neurons in population bursts of 2D living neural networks. New J Phys 10. doi:10.1088/1367-2630/10/1/015011. Yang H, Shew WL, Roy R, Plenz D (2012) Maximal variability of phase synchrony in cortical networks with neuronal avalanches. J Neurosci 32(3):1061–72. Johnston D, Brown T (1981) Giant synaptic potential hypothesis for epileptiform activity. Science (80- ) 211:294– 297. Bush PC, Douglas RJ (1991) Synchronization of bursting action potential discharge in a model network of neocortical neurons. Neural Comput 3(1):19–30. 12. Pinsky PF, Rinzel J (1994) Intrinsic and network rhythmogenesis in a reduced Traub model for CA3 neurons. J 8 Comput Neurosci 1(1–2):39–60. 13. Li C-YT, Poo M-M, Dan Y (2009) Burst spiking of a single cortical neuron modifies global brain state. Science (80- ) 324(5927):643–6. 14. Miles R, Wong RK (1983) Single neurones can initiate synchronized population discharge in the hippocampus. Nature 306(5941):371–373. 15. Takahashi H, et al. (2012) Light-addressed single-neuron stimulation in dissociated neuronal cultures with sparse expression of ChR2. BioSystems 107(2):106–112. 16. Okun M, et al. (2015) Diverse coupling of neurons to populations in sensory cortex. Nature 521:511–515. 17. 18. 19. Nowotny T, Rabinovich M (2007) Dynamical origin of independent spiking and bursting activity in neural microcircuits. Phys Rev Lett 98(12):128106:1-4. Beggs JM, Plenz D (2004) Neuronal avalanches are diverse and precise activity patterns that are stable for many hours in cortical slice cultures. J Neurosci 24(22):5216–29. Bakkum DJ, et al. (2013) Tracking axonal action potential propagation on a high-density microelectrode array across hundreds of sites. Nat Commun 4:2181. 20. Bakkum DJ, et al. (2013) Parameters for burst detection. Front Comput Neurosci 7(January):193:1-12. 21. Wagenaar D, Demarse TB, Potter SM (2005) MeaBench: A toolset for multi-electrode data acquisition and on-line analysis. 2nd Int IEEE EMBS Conf Neural Eng 2005:518–521. 22. Takens F (1981) Detecting strange attractors in fluid turbulence. Dynamical Systems and Turbulence, Lecture Notes in Mathematics, eds Rand DA, Young L-S (Springer-Verlag, Berlin), pp 366–381. 23. Sauer T, Yorke JA, Casdagli M (1991) Embedology. J Stat Phys 65(3/4):579–616. 24. 25. Sugihara G, May R (1990) Nonlinear forecasting as a way of distinguishing chaos from measurement error in time series. Nature 344(6268):734–741. Tsonis AA, Elsner JB (1992) Nonlinear prediction as a way of distinguishing chaos from random fractal sequences. Nature 358:217–220. 26. Sugihara G, et al. (2012) Detecting causality in complex ecosystems. Science (80- ) 338(6106):496–500. 27. 28. 29. Tajima S, Yanagawa T, Fujii N, Toyoizumi T (2015) Untangling brain-wide dynamics in consciousness by cross- embedding. PLOS Comput Biol 11(11):e1004537. Orlandi JG, Soriano J, Alvarez-Lacalle E, Teller S, Casademunt J (2013) Noise focusing and the emergence of coherent activity in neuronal cultures. Nat Phys 9(9):582–590. Shew WL, Yang H, Yu S, Roy R, Plenz D (2011) Information capacity and transmission are maximized in balanced cortical networks with neuronal avalanches. J Neurosci 31(1):55–63. 30. Hahn G, et al. (2010) Neuronal avalanches in spontaneous activity in vivo. J Neurophysiol 104(6):3312–22. 31. Shriki O, et al. (2013) Neuronal avalanches in the resting MEG of the human brain. J Neurosci 33(16):7079–90. 32. Beggs JM, Plenz D (2003) Neuronal Avalanches in Neocortical Circuits. J Neurosci 23(35):11167–11177. 33. 34. 35. 36. 37. Pinto DJ, Patrick SL, Huang WC, Connors BW (2005) Initiation, propagation, and termination of epileptiform activity in rodent neocortex in vitro involve distinct mechanisms. J Neurosci 25(36):8131–40. Ikegaya Y, et al. (2004) Synfire chains and cortical songs: temporal modules of cortical activity. Science (80- ) 304(5670):559–64. Luczak A, Barthó P, Harris KD (2009) Spontaneous events outline the realm of possible sensory responses in neocortical populations. Neuron 62(3):413–25. Eytan D, Marom S (2006) Cellular/Molecular Dynamics and Effective Topology Underlying Synchronization in Networks of Cortical Neurons. J Neurosci 26(33):8465–8476. Elger CE, Lehnertz K (1998) Seizure prediction by non-linear time series analysis of brain electrical activity. Eur J Neurosci 10(1997):786–789. 38. Golder S (2010) Tweet , Tweet , Retweet : Conversational Aspects of Retweeting on Twitter. 39. 40. Ishii A, et al. (2012) The "hit" phenomenon: a mathematical model of human dynamics interactions as a stochastic process. New J Phys 14(6):63018. Zhou J, Liu Z, Li B (2007) Influence of network structure on rumor propagation. Phys Lett Sect A Gen At Solid State Phys 368(6):458–463. 41. Gross T, D'Lima CJD, Blasius B (2006) Epidemic dynamics on an adaptive network. Phys Rev Lett 96(20):1–4. 9 42. 43. Pastor-Satorras R, Vespignani A (2001) Epidemic spreading in scale-free networks. Phys Rev Lett 86(14):3200– 3203. Burns SP, et al. (2014) Network dynamics of the brain and influence of the epileptic seizure onset zone. Proc Natl Acad Sci U S A. doi:10.1073/pnas.1401752111. 44. Wang Y, Goodfellow M, Taylor PN, Baier G (2012) Phase space approach for modeling of epileptic dynamics. Phys Rev E 85(6):61918. 45. 46. 47. 48. 49. 50. Stam CJ (2005) Nonlinear dynamical analysis of EEG and MEG: review of an emerging field. Clin Neurophysiol 116(10):2266–301. Lehnertz K, Elger CE (1995) Spatio-temporal dynamics of the primary epileptogenic area in temporal lobe epilepsy characterized by neuronal complexity loss. Electroencephalogr Clin Neurophysiol 95(2):108–17. Frey U, et al. (2010) Switch-matrix-based high-density microelectrode array in CMOS technology. IEEE J Solid- State Circuits 45(2):467–482. Livi P, Heer F, Frey U, Bakkum DJ, Hierlemann A (2010) Compact voltage and current stimulation buffer for high-density microelectrode arrays. IEEE Trans Biomed Circuits Syst 4(6 PART 1):372–378. Lorenz EN (1969) Atmospheric predictability as revealed by naturally occurring analogues. J Atmos Sci 26:636– 646. Kennel M, Abarbanel H (2002) False neighbors and false strands: A reliable minimum embedding dimension algorithm. Phys Rev E 66(2):26209. 10 Figure 1. Global bursts in cultured neural populations. (a) CMOS-based recording apparatus. (b) Spatial distribution of cells recorded in a representative preparation, Chip 1440. (c) Spontaneous activity of neurons in the same preparation, Chip 1440. The color indicates the spike count within each 5- ms time bin (white: 0 spikes; black >5 spikes). The abscissa represents the time. The majority of neurons fired together within network bursts (columns of dark-colored bins). (d) The temporal evolution of global mean field (normalized population firing rate) computed from c. The horizontal dashed line indicates the threshold for burst detection for this preparation. (Inset) The distributions of inter-spike intervals (ISIs) in five different preparations (solid lines) and the burst detection thresholds for the individual preparations (dashed lines). The horizontal axis shows the normalized log-ISI (the logarithm of inverse global-activity; see Methods). 11 Figure 2. Trace-back embedding (TBE) method in delay coordinates relates the global population activity to the local activities. (a) Delay reconstruction of the mean-field dynamics ("global states"). (b, c) Delay reconstructions of single neuron activities ("local states"). Two representative neurons (Cells 1 and 2) are shown. Blue dots: the time points of peak global activities during the bursts; Red dots: the temporally preceding states from the individual global activity peaks (here, shift time = -100 ms); magenta dots: the neighboring states within the reconstructed state space for Cell 1 (10 nearest neighbors for each burst peak); cyan pluses: predicted states based on TBE method. The protocol of TBE is as follows: (i) selecting a time point t on a local state trajectory (say, of neuron i) that corresponds to a peak population activity found in the global state trajectory (which is an target bursting state to be predicted), (ii) tracing-back the local state trajectory for a given time span, Δt, (iii) collecting "neighbor" time-points corresponding to states nearby t-Δt in the local state trajectory (avoiding temporally too near points), (iv) mapping the neighbor time-points back into the global state trajectory, and (v) proceeding the time with Δt on the global state trajectory to examine whether the finally obtained states (cyan pluses) cluster within bursting state and how accurately they replicate the target bursting state. The accuracy of burst prediction is higher when the traced-back states (red dotes) deviates more from the ones in non-bursting period in the local trajectory. Note that the tracing time in the retrograde direction with Δt (step ii) and mutual mapping between global and local states (steps i and iv) are newly introduced features in this study, which extends the previous methods (24–26). 12 a Figure 3. Presages of global burst within single neurons. (a) Burst detection with the TBE method based on a single representative neuron. The black traces represent global states in individual trials around burst peaks. The blue curve is the burst detection threshold for each detection time-span, Δt; global state exceeding this threshold provides burst predictor (or hallmarks of bursts after their occurrence). The threshold (dashed green line) was defined by 95 percentile of global activity outside bursting period. The red curve shows the fraction ("hit rate") of successfully detected bursts in all the burst trials. (b) Burst detection with TBE method based on the mean-field activity in the population. (c) Burst detection with the momentary firing rate of the same representative single neuron as the one shown in panel a. (d) Burst detection with the momentary mean-field firing-rate. (e) Burst predictability is a robust property of individual neurons. Dividing the data into two halves (the odd and even trials of spontaneous burst) shows that variability in burst-detection success rate is highly consistent over time. Each marker corresponds to a neuron. The filled markers represent the results using the mean field activity. The success rate was averaged over the range of time span -200 ms < Δt < -50 ms. (f) Burst detection with TBE method is more accurate than that based on the momentary firing rate. The comparison of success rates in burst detection with the TBE method and with the firing-rate-based method. Inset: the control analysis in which we compared the success rates of the TBE method with that of the firing-rate-based method using multiple time bins, by matching the number of bins to that used in the TBE method. The gray bars show the fraction of cells in all the preparations. The arrowheads indicate the medians. The colored lines show the fractions in the individual preparations. (g) Single cell can outperform the mean field at predicting the spontaneous network bursts, particularly when we utilize the information in the temporal patterns of neural activity rather than the momentary activity. The light and dark blue curves respectively show the TBE- and firing-rate-based success rates in burst detection with single neurons relative to that with the mean field. For each method, the cells were sorted based on the success rate. 13 Figure 4. The network structures in non-bursting periods explains the local burst predictability. (a) Causal network analysis. The network depicts the directed pairwise interactions among neurons. The filled (orange) and open (blue) markers represents the putative excitatory and inhibitory cells, respectively. The thickness of arrows indicates the absolute strength of causal coupling ("causality"). For clarity of illustration, only the top 5% strong couplings are shown, and the nodes are distributed by multi-dimensional scaling with the distances defined to be inversely proportional to the causality. The figure shows the result for Chip 1440. (b) Synaptically connected neuron pairs show larger causal interaction. The bar labels show the time span to be used for the causality analysis. Bars in dark and light blue: average cross-embedding values for cells that showed short-latent (dark, latency < 10 ms) and long (dark, latency > 10 ms) stimulus-triggered spike increase. The average causality index within time span [-100, 100] ms dissociated the post-synaptic cells from other cells (P<0.005, Wilcoxon sign rank test, independent samples). The error bar shows the s.e.m across neurons. (c) Identification of synaptic connectivity by electrical stimulation experiment. The figure shows the peri-stimulus time histograms of three representative neurons (from top to bottom, putative direct, indirect, and non-post cells, respectively). The asterisks indicate the earliest significant increase (P>0.05) of spike count, compared to the spike count distribution in the pre-stimulus period (from -500 to -100 ms). The figure shows the spike histogram smoothed with a boxcar kernel of 5- ms width for visualization purpose. (d) The relationship between burst predictability and the interaction strength in different neuron types (putative excitatory and inhibitory neurons). The neurons were classified into two groups depending on the strength of input or output causal couplings: the top half of neurons was labeled as "strong" whereas the bottom half was labeled as "weak." The error bar shows the s.e.m across neurons. (e) The same as panel d, but shows the individual neurons (the dots) with a 3D plot. The surfaces represent the fitted piecewise linear functions for each neuron types. The color conventions follow that of panel d. 14 Supplementary Figure S1. The population activity dynamics visualized with principal component analysis (PCA). The trajectories of population activity were plotted within the space of top three principal components (PCs 1–3), by applying PCA to the multi-neuron time series for each preparation. The results for three representative preparations are shown. 15 Supplementary Figure S2. The delay-embedding theorem (known as Takens' theorem). Suppose that we have the time-series of three variables, (𝑥(𝑡), 𝑦(𝑡), 𝑧(𝑡)) , generated by some differential equations: 𝑥 (𝑡) = 𝑓(𝑥(𝑡), 𝑦(𝑡), 𝑧(𝑡)), 𝑦 (𝑡) = 𝑔(𝑥(𝑡), 𝑦(𝑡), 𝑧(𝑡)) and 𝑧(𝑡) = ℎ(𝑥(𝑡), 𝑦(𝑡), 𝑧(𝑡)). (a) The evolution of the global state (𝑥(𝑡), 𝑦(𝑡), 𝑧(𝑡)) is represented by a trajectory in the space of those three variables. (b) Consider observing the temporal sequence of a single variable, e.g., 𝑥(𝑡). (c) The attractor topology in the original global state space is fully recovered in a delay-coordinate of the observed variable, (𝑥(𝑡), 𝑥(𝑡 − 𝜏), 𝑥(𝑡 − 𝜏)) with an arbitrary unit delay, 𝜏. Namely, we can construct a smooth one-to-one map from the reconstructed attractor to the original attractor. 16 Supplementary Figure S3. Identification of Excitatory and Inhibitory Neurons (a) Representative neurons in immunostaining with MAP2 and GABA. Action potentials below the insets were obtained at white rectangles, putatively from a neighboring neuron in a circle. The peak-to-peak time, Tpp, was defined as time duration from negative peak to positive peak of action potential. (b) Histogram of Tpp. Excitatory neurons (green) had larger Tpp than inhibitory neurons (magenta). K-means method to Tpp was employed to separate excitatory and inhibitory neurons. 17 Supplementary Figure S4. Microstimulation-based estimation of synaptic connectivity in a pair-wise manner. (a) Axonal stimulation on an arbitrary neuron elicited bidirectional action potential propagation. (b) Raw data of neural responses at a putative pre-synaptic neuron and post-synaptic neuron. Data from one hundred trials are superimposed. (c) Raster plot of (b). Antidromic, direct action potentials exhibited precise temporal response, while orthodromic, synaptic action potentials were elicited stochastically with significant temporal jitters. (d) Post-stimulus spike histograms of (c) 18
1708.00540
1
1708
2017-08-01T22:42:32
A silicon-based microelectrode array with a microdrive for monitoring brainstem regions of freely moving rats
[ "q-bio.NC" ]
Objective. Exploring neural activity behind synchronization and time locking in brain circuits is one of the most important tasks in neuroscience. Our goal was to design and characterize a microelectrode array (MEA) system specifically for obtaining in vivo extracellular recordings from three deep-brain areas of freely moving rats, simultaneously. The target areas, the deep mesencephalic reticular-, pedunculopontine tegmental- and pontine reticular nuclei are related to the regulation of sleep-wake cycles. Approach. The three targeted nuclei are collinear, therefore a single-shank MEA was designed in order to contact them. The silicon-based device was equipped with 3*4 recording sites, located according to the geometry of the brain regions. Furthermore, a microdrive was developed to allow fine actuation and post-implantation relocation of the probe. The probe was attached to a rigid printed circuit board, which was fastened to the microdrive. A flexible cable was designed in order to provide not only electronic connection between the probe and the amplifier system, but sufficient freedom for the movements of the probe as well. Main results. The microdrive was stable enough to allow precise electrode targeting into the tissue via a single track. The microelectrodes on the probe were suitable for recording neural activity from the three targeted brainstem areas. Significance. The system offers a robust solution to provide long-term interface between an array of precisely defined microelectrodes and deep-brain areas of a behaving rodent. The microdrive allowed us to fine-tune the probe location and easily scan through the regions of interest.
q-bio.NC
q-bio
A robust silicon-based microelectrode array, mounted on a microdrive A silicon-based microelectrode array with a microdrive for monitoring brainstem regions of freely moving rats G Márton1,2, P Baracskay3, B Cseri3, B Plósz4, G Juhász3, Z Fekete2,5, A Pongrácz2,5 1 Comparative Psychophysiology Department, Institute of Cognitive Neuroscience and Physiology, Research Centre for Natural Sciences, Hungarian Academy of Sciences, 2 Magyar Tudósok Blvd., H-1117, Budapest, Hungary 2 MEMS Laboratory, Institute for Technical Physics and Materials Science, Centre for Energy Research, Hungarian Academy of Sciences, 29-33 Konkoly Thege Miklós st., H-1121, Budapest, Hungary Research Group of Proteomics, Institute of Biology, Faculty of Science, Eötvös Loránd University, 1/C Pázmány P. Walkway, H-1117, Budapest, Hungary Plósz Microengineering Office Ltd., 119 Üllői st., H-1091, Budapest, Hungary MTA EK NAP B Research Group for Implantable Microsystems, 29-33 Konkoly Thege Miklós st., H-1121, Budapest, Hungary 3 4 5 E-mail: [email protected] 1 A robust silicon-based microelectrode array, mounted on a microdrive Abstract Objective. Exploring neural activity behind synchronization and time locking in brain circuits is one of the most important tasks in neuroscience. Our goal was to design and characterize a microelectrode array (MEA) system specifically for obtaining in vivo extracellular recordings from three deep-brain areas of freely moving rats, simultaneously. The target areas, the deep mesencephalic reticular-, pedunculopontine tegmental- and pontine reticular nuclei are related to the regulation of sleep-wake cycles. Approach. The three targeted nuclei are collinear, therefore a single-shank MEA was designed in order to contact them. The silicon-based device was equipped with 3×4 recording sites, located according to the geometry of the brain regions. Furthermore, a microdrive was developed to allow fine actuation and post-implantation relocation of the probe. The probe was attached to a rigid printed circuit board, which was fastened to the microdrive. A flexible cable was designed in order to provide not only electronic connection between the probe and the amplifier system, but sufficient freedom for the movements of the probe as well. Main results. The microdrive was stable enough to allow precise electrode targeting into the tissue via a single track. The microelectrodes on the probe were suitable for recording neural activity from the three targeted brainstem areas. Significance. The system offers a robust solution to provide long-term interface between an array of precisely defined microelectrodes and deep-brain areas of a behaving rodent. The microdrive allowed us to fine-tune the probe location and easily scan through the regions of interest. Keywords: Brainstem reticular formation; Freely moving rats; Single unit recording; Silicon-based probe; extracellular microelectrode array; neural; microdrive 1. Introduction Recording electrical activity in different brain areas simultaneously is increasingly attracting the attention of the neuroscience community [1-3], which creates new approaches in the system design of implantable devices as well. One of the main aims of researchers is to record as many neurons as possible from different brain areas, searching for synchrony and time-locking of activities. The applicable number of electrodes is limited by the tissue injury made by a certain electrode construction and the size and distribution of recording sites. Up to now, plenty of different electrode designs have been introduced such as multiple microwire arrays or bundles [4-7], micromachined silicon-based microelectrode arrays (MEAs) [8], polymer- [9], ceramic-based [10] and titanium [11] probes. When extracellular electrodes are chronically used, signal detection is hindered by glial scar formation around the implants [12, 13]. This is one of the reasons why it is advantageous if the implanted probes can be precisely moved postoperatively. This function can be enabled with the use of microdrives [14, 15]. Modern silicon-based microtechnology, fuelled by the electronic industry allows precise batch fabrication of microelectromechanical systems (MEMS) with reproducible features and small dimensions, including versatile and cost-effective ways of making neural MEAs [16]. Numerous electrodes can be placed at different points along each probe, moreover, they offer the potential of integrating signal processing circuits on them. Silicon, along with the usually applied silicon dioxide 2 A robust silicon-based microelectrode array, mounted on a microdrive and silicon nitride insulator thin films [17], is a biocompatible material [18, 19]. Silicon can be tailored e.g. with deep reactive ion etching (DRIE) in order to form versatile probe geometries [20]. Thin-film microelectrodes, located on the silicon probes are suitable for in vivo local field potential and unit recordings [21-23]. They are widely used in freely moving animals for measurements from different brain regions, such as the neocortex or the hippocampus [24, 25]. Silicon-based microtechnology allows one to equip neural probes with delivery channels for injecting chemical compounds close to the active recording sites [26-28] or optical waveguides for optogenetic experiments [29]. According to these considerations, silicon is an excellent candidate to be substrate material for neural MEAs. However, mechanical properties of silicon microstructures are not always satisfactory. Not only the shafts of the probes can break, but cracking and degradation of the electric wires can occur. Such failures were detected in 15 µm thick MEAs, chronically implanted into mice [30]. Failures due to poor mechanical properties might be the main reason why silicon-based MEAs are hardly used in deep-brain regions even in studies involving rodents (with some exceptions [31]), while mechanical strains can be larger in case of higher mammals [30]. Recent studies revealed that silicon probes with "ultra-long" shafts can be successfully fabricated [32, 33], and these reinforced or relatively thick devices might be able to withstand larger forces [34]. The wide-spread application of ultra-long" silicon-based MEAs for discovering deep-brain regions might contribute to the progress of neurosciences. Our neuroscientific aim was the simultaneous recording of neural activity from three nuclei of the brainstem reticular formation: the deep mesencephalic reticular nucleus (DpMe), the pedunculopontine tegmental nucleus (PPTg) and the oral part of the pontine reticular nucleus (PnO). The targeted structures are part of the ascending reticular activating system (ARAS) and play a key role in the regulation of different functional brain states such as the sleep-wake cycle, hypersynchronization, desynchronization and arousal through integrating and transmitting different sensory inputs from the spinal cord and information from other brainstem areas [35, 36]. The ARAS consists of neuron-poor nuclei and a gigantocellular neural networks, called the reticular formation. Although these nuclei seem separate, many of them are overlapping and they are technically one anatomical unit. These areas give ascending projections to the cortex, the basal ganglia, and the thalamus, ascending descending connections to the spinal chord, and they are richly interconnected with each other [37]. Selective lesion of the PnO induces chronic comatic state, thus the PnO likely acts as an activating centre [36]. Because of the rich anatomical interconnections [38, 39], investigating functional correlation of neuronal activity from these three nuclei is an intriguing question. Extracellular neural activity of the selected areas has separately been measured in head-restrained or anaesthetized animals [40-44]. Nuclei of the ascending reticular activating system were targeted by a number of experiments using microwires in freely moving cats and rats [43, 45, 46]. However, these results are restricted to one area at a time. To the authors' knowledge, no simultaneous extracellular unit recordings from several reticular formation areas in freely moving rats have been presented. Insulated wire electrodes are frequently employed for recordings from other deep brain areas as well [47]. In contrast with these techniques, we have decided to create a probe utilizing the above described silicon-based MEMS technology, in order to interface with the deep brain structures in our interest. We expected that the silicon shaft would allow sufficient, more precise targeting than microwires, moreover, by implanting a specific formation of custom-designed array of electrodes at once, instead of managing individual wire electrodes or bundles of them, time and effort could be saved. The brain structures of interest are collinear, so we decided to create a single-shaft MEA with electrodes custom- 3 A robust silicon-based microelectrode array, mounted on a microdrive arranged into three groups. We also aimed to construct a microdrive in order to finely relocate the sensor and find spots for successful electrophysiological recordings, containing unit activities. Another trivial way of interfacing with the three nuclei would have been the utilization three individual MEAs. Mounting a microdrive system on the skull of a rat, which is suitable for handling the shafts of three probes is probably a challenging, yet not impossible task. That would have allowed more freedom in positioning, but caused more tissue damage and cost more preparation time and resources. In summary, this work focuses on a system, with the help of which simultaneous recording of unit activities of three deep brain structures was successfully performed in behaving rats for the first time. It will describe the measurement methods, the neurobiological analysis of the obtained results exceeds its scope. 2. Materials and methods 2.1. Probe design A series of masks have been designed to fit the electrode array geometry to the anatomy of the brain regions we proposed to investigate. Accordingly, the total length of the probe is 30 mm, the shaft length is 25 mm, and the shaft thickness is the 200 μm initial thickness of the Si-wafer. The width of the probe gradually increases from 60 μm up to 130 μm along the recording sites, and up to 400 μm at the base of the probe to keep the robustness of the design. The distance between the tip and the middle of the first recording site is 92 μm. The tip angle is 19°. This design contains square-shaped (12 m x 12 m) platinum recording electrodes in three groups, located up to 3.2 mm from the probe tip. One group consists of 4 electrodes with 150 µm site spacing, and the groups are 780 µm distant from each other. The output platinum leads are 4 µm wide and 300 nm thick. 2.2. Microfabrication The fabrication technology of the multielectrodes is based on a single-side, three-mask bulk micromachining process performed in three phases: (a) bottom passivation layer deposition and lift-off process for the formation of platinum electrodes, output leads and bonding pads, (b) top passivation layer deposition and patterning, (c) etching of the passivation layers and (d) deep reactive ion etching (DRIE) of silicon to form the probe body. The overall process flow is summarized in figure 1. 4 A robust silicon-based microelectrode array, mounted on a microdrive Figure 1. Cross-sectional schematic diagram of the main fabrication steps. (a) SiO2 and SixNy layer deposition and patterning of platinum wiring. (b) Deposition and selective opening of the SixNy passivation layer. (c) Patterning of front-side Al mask and dry etching of the passivation and insulation layers; evaporation of Al and spin-coating of photoresist to form the backside etch-stop layer. (d) Cross- section of the ready-made probe after removing the masking and etch-stop layers. 2.2.1. Electrode array formation 200 μm thick, double-side polished (100) oriented 4-inch silicon wafers are used for probe fabrication. In the first step a 100 nm thick SiO2 layer was thermally grown, followed by deposition of a 500 nm thick LPCVD low-stress silicon nitride. The metal layer was then deposited and patterned by lift-off process. The lift-off structure consisted of 1.8 μm thick photoresist (Microposit 1818) layer over patterned 500 nm thick Al thin film. The metal layer consisted of a 15 nm thick adhesion layer of TiOx and Pt. TiOx was formed by reactive sputtering of Ti in O2 (Ar/O2 ratio was 80:20) atmosphere. In the same vacuum cycle 270 nm thick Pt was sputtered on top of TiOx. The lift-off was accomplished by dissolving the photoresist pattern in acetone, this process was optimized by using water-cooled substrate holder which diminished the resist distortion during TiOx/Pt sputtering. Subsequently, the removal of Al patterns in a fourcomponent etching solution and low pressure chemical vapor deposition of 300 nm thick SiNx insulating layer stack were performed. Contact holes were etched through the SiNx layer to unblock the electrical recording sites at the probe tip and the pads at the base for wire bonding. 2.2.2 Probe shaping A 500 nm thick Al layer was deposited on the front side and patterned by photolithography using 4 µm thick SPR220 photoresist. The contour of the probe body was defined by Al wet etching. SPR220 photoresist was spun on the backside of the wafer acting as a stopping layer during the subsequent deep reactive ion etching of silicon. The 3D micromachining process was performed in an Oxford Plasmalab System 100 chamber using Bosch process. 2.3. Packaging Rigid-flexible PCBs have been formed in order to provide sufficient support for the silicon probes and enough freedom for the movements of the microdrive. The deformability of this component is illustrated in figure 2. The substrate of its rigid parts was made of 1-mm thick FR-4. The substrate of the flexible cable was manufactured from 50 µm thick polyimide. A conductor layer of copper has been formed on 5 A robust silicon-based microelectrode array, mounted on a microdrive the top surface, plated with gold for the ease of wire bonding. The upper insulator layer on the PCB was made of flexible solder-resist mask. The geometry of the device has been designed as follows. A 6 mm x 9 mm rigid part was shaped for the support of the probe. A hole of 2 mm in diameter was designed for fixing the PCB with a screw onto microdrive. The L-shaped flexible cable was 5 mm wide and 19.1 mm long at its midline. It provided electronic connection via 200-µm thick wires between the probe support and the Preci-Dip connector part of the PCB. The connector part was rigid, suitable for soldering a 12-pin section of a 90-degree angle pin connector (Preci-Dip 850-PP-N-050-20-001101). Figure 2. The rigid-flexible PCB, containing a rigid part for probe support (1), the flexible polyimide cable (2) and the Preci-Dip connector part (3). 2.4. Microdrive The extended length of the probe requires stable and sufficiently vibration-free system to minimize tissue damage along the probe track, while providing flexibility to move the implanted electrode shank in the brain. A novel microdrive configuration was designed and is illustrated in figure 3. The 8 mm wide, 7 mm thick, 25 mm high, body of the device weights 2.7 g and consists of three, half octagonal prism-shaped blocks. The upper block (5), which contains the thread of the driving screw (5), and the lower block (7) are made of aluminium. The electrode (4) can be fastened through the rigid part of the PCB (3) to the microdrive by a small screw (9). The middle, electrode-carrier block (6) is held by two tightly fit rods (8) and is driven by the large driving/positioning screw (10). This block is made of polytetrafluorethylene (PTFE), as this self-lubricating type plastic renders the opportunity to highly reduce the joint gap between the block and the rods. Thus vibration is reduced to the minimum, resulting in a really robust and stable system. One 360° turn of the driving screw moves the probe tip with 300 µm, without any rotational movement. 6 A robust silicon-based microelectrode array, mounted on a microdrive Figure 3. Layout of the Si-electrode - microdrive system. (a) Front (left) and rear (right) view of the assembled system. 1: connector; 2: flexible cable; 3: the rigid part of the PCB; 4: the probe shaft; 5: microdrive upper block with the thread for the driving screw; 6: microdrive middle, electrode-carrier block; 7: microdrive lower block; 8: two tightly fit rods; 9: fastening screw; 10: driving screw. (b) Assembling of the driving screw, the microdrive and the electrode via the fastening screw. (c) Schematic drawing of the sensor region of the shaft, with three groups of sites. 2.5. Electrochemical impedance spectroscopy In order to characterize the electrodes, electrochemical impedance spectroscopy (EIS) measurements were performed in physiological saline, using an Ag/AgCl reference electrode (Radelkis Ltd., Hungary) and a counter electrode of platinum wire with relatively high surface area. The probe signal was sinusoidal, 25 mV RMS. A Reference 600 instrument (Gamry Instruments, PA, USA) was used as potentiostat and Gamry Framework 6.02 and Echem Analyst 6.02 software were used for experimental control, data collection and analysis. Experiments were performed in a Faraday cage. 2.6. Animal surgery Eleven adult Sprague-Dawley rats (Charles River Laboratories, from the breeding colony of ELU, Hungary) were implanted with the probes under 4% isoflurane anaesthesia. The care and treatment of all animals were conformed to Council Directive 86/609/EEC, the Hungarian Act of Animal Care and Experimentation (1998, XXVIII), and local regulations for the care and use of animals in research. All efforts were taken to reduce the number of animals used and to minimize the animals' pain and suffering. Prior to surgery we chronically applied the antibiotic Enorofloxacin in 10 mg/kg dose to prevent inflammatory reactions. We mounted the animals onto a stereotaxic instrument (David Kopf Instruments, California, USA), and a longitudinal cut was made on the scalp. Skin and muscles were retracted and the surface was cleaned. Four screw electrodes over the cortex were used to record the electrocorticogram (ECoG): one over the right secondary motor cortex (AP: 4.5 mm, ML: 2.5 mm with reference to the bregma), one over the right somatosensory cortex (AP: -2.9 mm, ML: 2.5 mm), two 7 A robust silicon-based microelectrode array, mounted on a microdrive over the right and left visual cortex (AP: -8,3 mm, ML: 2,5 mm). Two additional screw electrodes over the cerebellum served as grounding. After driving the screw electrodes into the cranium, the skull was opened over the left hemisphere using a drill, and the dura mater was incised. The neural probe was fastened to the microdrive that was adjusted to the stereotaxic instrument through a custom-made adaptor, and the probe was inserted into the brain manually by oblique targeting (47° inclination to the vertical), which made it possible to span all three target areas through a single electrode track at the same time. After the initial penetration, silicone-vacuum grease (Beckman Coulter Inc., Fullerton, CA, USA) was applied on the probe shank at the brain surface to provide smooth penetration further into the tissue and prevent blood clot formation. Finally, SuperBond resin cement (Sun Medical Company Ltd., Moriyama, Japan) was applied on the skull surface and the microdrive was fixed. A crown-like structure was formed around the system, utilizing UniFast dentacrylic cement (GC America Inc., Chicago, USA) and sterile plastic components fabricated from syringes and Petri dishes, protecting its external components. The structure typically weighted 12 g, was 28 mm long, 22 mm wide, and 30 mm high. The microdrive – and thus the implant - was not sealed hermetically from the environment, therefore proper hygienic conditions were in focus to avoid infection at the craniotomy. The implantation is illustrated in figure 4. 2.7. Course of experiments During recordings, the rat was kept in a cage of 40 cm × 60 cm, surrounded by a Faraday cage for electrical shielding. The connecting cable was continuously held up by a light mechanical lever, which prevented the rat from reaching the cable and damaging it, without impeding the animal in its movements. Nine rats were allocated for gaining high amount of data considering the characteristics of the targeted brain nuclei. To achieve this, we needed short, 1-2 hour sessions, during which the rats changed from awakeness to sleeping state or vice versa. During these sessions, we measured firing patterns of individual cells. After a successful session (or on the following day) we used the microdrive to relocate the probe so that new groups of cells could be explored. During the initial implantation, the probe tip was aimed at the area of the pedunculopontine tegmental nucleus. The Bregma was used as reference point for targeting, Instead of surface coordinates, a depth coordinate was determined (AP -7.0 ; ML 1.5; dorsoventral 7.6), to which the tip was targeted in a 47° angle and from where the oblique path down to deeper areas started. The tilting of the head in the sagittal plane was avoided by maintaining the "flat skull position" by checking and refining the correctness of the lambda (DV = 0) and interaural line (AP = 9.0; DV = 10.1) coordinates in reference to the bregma. During insertion, the probe was slowly (approximately 2-3 mm/min), manually advanced, solely with the usually dorsovenrtral directional drive of the stereotaxic apparatus (which was tilted so that the driving direction was parallel to the direction of the probe shaft). Thus each electrode group could record from different deep brain areas over time. At the deepest penetration the tip reached the caudal pontine reticular nucleus. Altogether, the system allowed us to record 12 unit channels (4 from each target area) and 4 ECoG channels from each animal. The electrode array was slowly advanced (60-100 µm per day) during the recording period (minimum 22, maximum 125 days) and stopped if extracellular unit firing coud be recorded, until the electode tip reached the PnC. Two additional rats were sacrificed in order to gain some information about the chronic static performance of the electrodes (without driving). In this case, we used the very same methodology for 8 A robust silicon-based microelectrode array, mounted on a microdrive implantation, as previously described, but did not drive the probe so frequently, as instead, we kept them at the same location for weeks. Figure 4. Implantation surgery and histological evaluation of electrode position. (a), (b) Positioning and inserting of the electrode into the brain tissue. (c) Implanted rat after surgery. (d) Illustration of the implantation strategy. Figure based on Fig. 82. of [48]. PnC: pontine reticular nucleus, caudal part; PnO: pontine reticular nucleus, oral part; PPTg: pedunculopontine tegmental nucleus 2.8. Recording system and data analysis methods We used a multichannel amplifier (Multiamp SMA-1a, Supertech) and a data processing system (CED Micro1401, Cambridge Electronics Design Ltd., Cambridge, UK) for data acquisition. Extracellular action potentials were captured at 33 kHz on each electrode whenever their amplitude reached the 100 µV threshold. We used a 500 – 8000 Hz online filter and a 300 – 6000 Hz offline filter before spike sorting. Electrocorticogram digitalization frequency was 5 kHz (offline filtering between 0.2 and 300 Hz). Data files were recorded with Spike2 (Cambridge Electronic Design Ltd., Cambridge, UK) software. Spike waveforms were sorted by using a 3D off-line spike analysis program (OFS/v2, Plexon Inc., Dallas, TX, USA). With the help of principal component analysis, we determined scatterplots using the first two principal components. K-means clustering analysis was used to define cluster borders of single units. ECoG spectrograms, autocorrelations and cross-correlations were calculated to validate 9 A robust silicon-based microelectrode array, mounted on a microdrive the cluster verification with NeuroExplorer (Plexon Inc., Dallas, TX, USA). To assess the quality of spike cluster separation, we investigated the spike contamination of the refractory period in the autocorrelograms [49]. If such contaminations were present, additional efforts, such as manual clustering were made to separate the waveforms. In order to quantify the quality of the unit cluster separation, we employed three statistical methods: a multivariate analysis of variance (MANOVA) test (F), a nonparametric test (J3) and the calculation of the Davies-Bouldin validity index (DB), which is also nonparametric. These methods are described in detail in the supporting text of a research paper by Nicolelis et al. [50]. The J3 index is calculated by dividing the average distance between clusters (J2) by the average distance of points from their cluster mean (J1). Therefore, higher J3 values indicate better separation rates. The DB index is defined as 𝐷𝐵 = 1 𝑛 𝑛 𝑘=1 ∑ max 𝑘≠𝑙 { 𝑆𝑛(𝑄𝑘)+𝑆𝑛(𝑄𝑙) 𝑑(𝑄𝑘,𝑄𝑙) } , (1) where n is the number of clusters, Sn(Q) is the average distances of the elements of cluster Q to the center of cluster Q and d(Qk,Ql) gives the distance between the center of the Qk and Ql clusters [50]. Good clustering is characterized by high F, J3 and small BD index values. 2.9. Histological evaluation Electrode placement can be verified by electrical stimulation through the recording sites in rats deeply anesthetized with urethane [51, 52]. We used 200–400 ms square wave current pulses with a current strength of 1–10 μA. Animals were then transcardially perfused with 4% paraformaldehyde in sodium- phosphate buffer. 60 µm longitudinal sections were cut from the brains using a freezing microtome, and the tissue was stained with the Gallyas silver-staining method [53, 54]. Thus the electrode track of the chronically inserted probe and acutely injured „dark" neurons caused by the electrical stimulation can be seen at the same time. 3. Results and discussion 3.1. The realized system The custom design technology allowed precise electrode array arrangement on the silicon probe, fitted to the location of the three brain areas we intended to observe. Photographs and scanning electron microscopic images of the MEA are shown in figure 5(a)-(c). The results of the EIS measurements are shown in figure 5(d). At 1 kHz, 3.12 MΩ ± 0.21 MΩ average impedance magnitude was yielded for the 12 × 12 µm2 electrodes, which is in consistence with the literature [55]. The robustness of the device has proven to be sufficient for implantation. We have observed neither breaking nor bending of the shafts. The flexible connection to the socket and firm holding of the electrode in a hard printed circuit board (PCB) attached to the microdrive kept the stability of the probe and provided freedom for driving. While custom designed microdrives are commonly used for the manipulation of fine wire electrodes, tetrodes or hexatrodes [56-58], it is less frequent in case of silicon-based MEAs. With a silicon-based MEMS thermal actuator, driving of a microelectrode array with 8.8 µm step resolution was achieved [59]. However, the construction of the device requires deposition of 11 layers onto the substrate wafer, including 4 layers of polysilicon. In our case, the driving was achieved without involving subtle motors 10 A robust silicon-based microelectrode array, mounted on a microdrive or active elements. Compared to the dDrive system of NeuroNexus (Ann Arbor, MI, USA), the system presented here utilizes aluminium blocks instead of plastic, and two vertical aligners improve stability of the vertically moving block rather than purely relying on the driving screw. The whole system met the particular demands of the angled implantation and deep location of the targeted areas. Figure 5. (a) The silicon probe mounted onto the rigid-flexible PCB. (b)-(c). Scanning electron microscopic images of the probe tip and a recording site. (d) Average EIS spectrum of the electrodes of a probe. Standard deviation values were less than 10% of the corresponding magnitude values on all frequencies. At 1 kHz, the magnitude of the impedance was 3.12 MΩ ± 0.21 MΩ. The microdrive was stable during the recording sessions. The stereotaxic instrument provided a specific targeting angle and precise implantation during surgery, and the microdrive allowed uni-directional positioning of the MEA even after six months. Figure 6 shows images of the histological findings. The dark neurons of the Gallyas staining method indicated an approximately 4-600 µm wide trail of perturbed tissue along the probe track and confirmed successful targeting of the deep-brain regions of interest. No sign of bending was observed on the used, explanted probes. 11 A robust silicon-based microelectrode array, mounted on a microdrive Figure 6. Electrode position evaluated with histology. Slices were stained with Gallyas method. Dark neurons indicate the position of the probe shaft in the brain tissue. (a)-(b) Electrode track with a schematic drawing of the inserted probe. Scale bars are respectively 1.5 mm and 0.8 mm. Abbreviations: 7n: facial nerve or its root; DpMe: deep mesencephalic nucleus; LC: locus coeruleus; lfp: longitudinal fasciculus of the pons; MVePC: medial vestibular nucleus, parvicellular part; MVPO: medioventral periolivary nucleus; Pn: pontine nuclei; PnC: pontine reticular nucleus, caudal part; PnO: pontine reticular nucleus, oral part; (c)-(e) Zoomed images of the probe track in the targeted nuclei. Scale bars are 400 µm. 3.2. Unit activities recorded from the targeted nuclei The majority of the implanted MEAs (n = 9) were utilized in order to obtain single unit activities from different cells in the nuclei of freely moving rats. As mentioned earlier, in section 2.7. Course of experiments, these probes were driven further with the microdrive after successful, 1-2 hour recording sessions, which contained transitions from awake/sleeping states. We have successfully recorded single unit activity chronically from all target areas (DpMe, PpTg, PnO). Figure 7 shows examples of recorded units during an epoch of 15 seconds in a rat, 25 days after surgery. Raw data and spectrogram of the ECoG, raw data and unit firing from 4 multi-unit recording sites and the averaged waveforms of sorted individual units are presented. The spectrogram of the ECoG reveal sleep spindles, starting at around t = 2 and 14 seconds. Spike trains of cells are presented that fire frequently at the beginning of the spindle (M1 units), and ones that cease to be active at the end of the spindle (B2, B3 units). The data gained by the MEAs gave insight into the firing patterns of different cells that take part in sleep regulation in the brainstem. We were able to record approximately 200 extracellular units from this animal, of which 163 were thoroughly analyzed after sorting. Units that fired for just a short period of time or had too low firing frequency or high spike waveform variance were excluded. Of the 163 neurons, 15 were present on a single electrode site only. Thus we recorded 148 neurons concurrently. The active contacts between 12 A robust silicon-based microelectrode array, mounted on a microdrive parallel recordings were minimally 2, maximally 6 at the same time. The contacts showed unit activity in ~27.5% of the cases with an average of 5 recorded cells. These results show that individual cell activities in deep brain regions of a rodent can be recorded with relatively high yield using such silicon probes. The number of simultaneous recordings of sorted units from the target nuclei (DpMe, PPTg, PnO) was achieved, however, sparsely: altogether 13 such sessions occurred in the 9 rats out of altogether 266 analyzed sessions (on the average, 1.4 ± 0.5 per rat). Figure 7. Representative parallel recordings from different sites on an implanted probe in a freely moving rat, 25 days after surgery. (a) Raw data and spectrogram of the ECoG channel, and unit channels with single unit patterns (spike trains). Elevated power in the lower frequency range of the ECoG channel indicates sleep spindles starting at around t = 2 and 14 seconds. Activities of cells that are in connection with the spindles were recorded. (b) 0.24 second long recording sections (highlighted with a red band on panel A) at a faster time-base after offline filtering (ECoG: 0.2 – 300 Hz, probe site channels: 300-6000 Hz). (c) Different types of representative autocorrelograms show heterogeneous spiking properties. 13 A robust silicon-based microelectrode array, mounted on a microdrive 3.3. Signal stability Two additional rats were implanted in order to gain insight into the static chronic performance of the electrodes. These MEAs were not driven after each successful recording session, as instead, we let them stay in the same location for 2-7 days. Figure 8 shows the functionality of such a probe during a 33-day period. The number of different single unit activities varied from 1 to 8, with an average of 3.9 ± 2.0. Figure 8(a) shows the number of single unit activities on each day, when recordings were performed. Black triangles mark the occasions when the probe was driven further into the tissue. The quality of unit clustering is shown in figure 8(b). The mean ± standard error for the indices throughout the 33-day period were the following: F = 8.23 ± 5.43, J3 = 2.76 ± 1.53, DB = 0.26 ± 0.1. For these values, the measurements of Nicolelis et al. with microwire arrays in the cortex of rhesus monkeys yielded F = 11.5 ± 1.0, J3 = 3.8 ± 0.4, DB = 0.32±0.01 [50]. Figure 8(c)-(c') illustrate the results of cluster analysis of three cells recorded from the same animal on day 21 and 22, respectively, showing that good quality single units of the same clusters can be recorded on different days. 14 A robust silicon-based microelectrode array, mounted on a microdrive Figure 8. (a) The number of single unit activities on each day, when recordings were performed from a rat. Black triangles mark the occasions when the probe was driven further into the tissue. (b) The quality of unit clustering. The mean ± standard error values considering the 33-day period: F = 8.23 ± 5.43, J3 = 2.76 ± 1.53, DB = 0.26 ± 0.1. (c)-(c') Typical unit clusters on a channel during a session and the same clusters obtained from the signals recorded on the following day. Horizontal and vertical scale bars on the spike waveforms are 0.5 ms and 100 µV, respectively. The cluster plots contain data of 120 minute long recordings. Unit activities on 20-30% of the electrodes on 15 µm thick planar Michigan MEAs were reported 30 days post-implant, in the motor cortex of rats [60]. With such probes, 10-15% unit yield was measured after 3-6 months, even though a significant portion of the electrodes was in white matter [13]. Comparing our results to those findings, the long-term performance of our system is slightly poorer. The causes of this can be various, we provide three possibilities here. Firstly, highly different brain 15 A robust silicon-based microelectrode array, mounted on a microdrive regions might respond differently to the implantation: the cells in our targeted nuclei might be more sensitive compared to the cells of the cortex or the hippocampus. Such heterogeneity in unit stability was even observed in different layers of the cortex: layers IV to VI were claimed to have the greatest yield acutely, but layers II to IV the best yield in chronic time points [13]. Secondly, the much larger (200 µm) shaft thickness of our probes is disadvantageous in this aspect, it can induce a more aggressive immune response [61]. Finally, interfacing with the deep brain nuclei required a much longer implantation track than the cortical areas would have required, which results in a bigger chance of disrupting blood vessels, thus facilitating chronic signal degradation [12]. It was not our goal to measure the same unit activities for weeks or months (which is a common priority e.g. for the development of brain-computer interfaces with invasive electrodes). Rather, we intended to observe firing patterns of different cells of deep brain nuclei during 1-2 hour sessions. In the light of our aims, the microdrive alleviated the problem of losing signals of unit activities, as slight motions of the probe could bring the electrodes into less perturbed tissue regions and explore new cells. 4. Conclusions We report here on a novel silicon-based MEA design, tailored to be used in a specific neuroscientific experiment, which required interfacing with three deep brain nuclei related to the regulation of sleep- awake cycles. The applied fabrication technology provided cost-effective batch production of the MEAs. The stiff body of the probe allowed sufficient implantation, even into the brainstem, while the custom-designed electrode array formation provided coverage of the investigated brain regions. The probe was suitable for measuring unit activities in the targeted structures, with high yield, through several recording sites of freely moving rats, fulfilling our goals. The main novelty of our system is the utilization of a special microdrive, which enabled us to accurately relocate the electrodes after the implantation, and explore new cells. Taking into account that the brainstem is one of the most difficult targets for neural recording in rats, our system can probably be used in many different experiments, when two or more brain structures can be exposed in an electrode track, even if the track is not perpendicular to the cerebral surface. In a future study, we intend to investigate the usability of such MEAs in the central nervous system of larger mammals, such as cats or primates. Acknowledgements We thank András Czurkó for critical reading of the manuscript. We are thankful for the laboratory personnel of the MEMS Laboratory of the Institute for Technical Physics and Materials Science. This work was supported by the KTIA_13_NAP-A-IV/1-4;6 research grants. Z. Fekete is grateful for the support of the Hungarian Brain Research Program (KTIA NAP 13-2-2014-0022). A. Pongracz is grateful for the Bolyai Janos scholarship of the Hungarian Academy of Sciences. References Fujisawa S and Buzsaki G 2011 A 4 Hz oscillation adaptively synchronizes prefrontal, VTA, and hippocampal activities Neuron 72 153-65 Buzsaki G 2010 Neural syntax: cell assemblies, synapsembles, and readers Neuron 68 362-85 16 [1] [2] A robust silicon-based microelectrode array, mounted on a microdrive [3] [4] Buzsaki G 2004 Large-scale recording of neuronal ensembles Nature neuroscience 7 446-51 Kubie J L 1984 A driveable bundle of microwires for collecting single-unit data from freely-moving rats Physiol Behav 32 115-8 [5] Wilson M A and McNaughton B L 1994 Reactivation of hippocampal ensemble [6] [7] memories during sleep Science 265 676-9 Rinaldi P, Juhasz G and Verzeano M 1976 Analysis of circulation of neuronal activity in the waking cortex Brain Res Bull 1 429-35 Verzeano M and Calma I 1954 Unit activity in spindle bursts Journal of neurophysiology 17 417-28 [8] Wise K D and Najafi K 1991 Microfabrication techniques for integrated sensors and [9] microsystems Science 254 1335-42 Lai H-Y, Liao L-D, Lin C-T, Hsu J-H, He X, Chen Y-Y, Chang J-Y, Chen H-F, Tsang S and Shih Y-Y I 2012 Design, simulation and experimental validation of a novel flexible neural probe for deep brain stimulation and multichannel recording Journal of Neural Engineering 9 036001 [10] Moxon K A, Leiser S C, Gerhardt G A, Barbee K A and Chapin J K 2004 Ceramic- based multisite electrode arrays for chronic single-neuron recording IEEE transactions on bio-medical engineering 51 647-56 [11] McCarthy P T, Rao M P and Otto K J 2011 Simultaneous recording of rat auditory cortex and thalamus via a titanium-based, microfabricated, microelectrode device Journal of neural engineering 8 046007 [12] Kozai T D, Jaquins-Gerstl A S, Vazquez A L, Michael A C and Cui X T 2015 Brain tissue responses to neural implants impact signal sensitivity and intervention strategies ACS chemical neuroscience 6 48-67 [13] Kozai T D, Du Z, Gugel Z V, Smith M A, Chase S M, Bodily L M, Caparosa E M, Friedlander R M and Cui X T 2015 Comprehensive chronic laminar single-unit, multi- unit, and local field potential recording performance with planar single shank electrode arrays J Neurosci Methods 242 15-40 [14] Otchy T M and Lveczky B P 2012 Design and assembly of an ultra-light motorized microdrive for chronic neural recordings in small animals J Vis Exp 8 [15] Vandecasteele M, M S, Royer S, Belluscio M, Berenyi A, Diba K, Fujisawa S, Grosmark A, Mao D, Mizuseki K, Patel J, Stark E, Sullivan D, Watson B and Buzsaki G 2012 Large-scale recording of neurons by movable silicon probes in behaving rodents J Vis Exp 4 [16] Fekete Z 2015 Recent advances in silicon-based neural microelectrodes and microsystems: a review Sensors and Actuators B: Chemical 215 300-15 [17] Drake K L, Wise K D, Farraye J, Anderson D J and BeMent S L 1988 Performance of planar multisite microprobes in recording extracellular single-unit intracortical activity Biomedical Engineering, IEEE Transactions on 35 719-32 [18] Schmidt S, Horch K and Normann R 1993 Biocompatibility of silicon-based electrode arrays implanted in feline cortical tissue J Biomed Mater Res 27 1393-9 [19] Neumann A, Reske T, Held M, Jahnke K, Ragoss C and Maier H R 2004 Comparative investigation of the biocompatibility of various silicon nitride ceramic qualities in vitro J Mater Sci Mater Med 15 1135-40 [20] Cheung K, Gun L, Djupsund K, Yang D and Lee L P 2000 A new neural probe using SOI wafers with topological interlocking mechanisms. In: Microtechnologies in Medicine and Biology, 1st Annual International, Conference On. 2000, pp 507-11 17 A robust silicon-based microelectrode array, mounted on a microdrive [21] Grand L, Wittner L, Herwik S, Gothelid E, Ruther P, Oscarsson S, Neves H, Dombovari B, Csercsa R, Karmos G and Ulbert I 2010 Short and long term biocompatibility of NeuroProbes silicon probes Journal of neuroscience methods 189 216-29 [22] Nicolelis M A and Ribeiro S 2002 Multielectrode recordings: the next steps Current Opinion in Neurobiology 12 602-6 [23] Szarowski D H, Andersen M D, Retterer S, Spence A J, Isaacson M, Craighead H G, Turner J N and Shain W 2003 Brain responses to micro-machined silicon devices Brain Research 983 23-35 [24] Csicsvári J, Henze D A, Jamieson B, Harris K D, Sirota A, Bartho P, Wise K D and Buzsáki G 2003 Massively parallel recording of unit and local field potentials with silicon-based electrodes Journal of neurophysiology 90 1314-23 [25] Bragin A, Hetke J, Wilson C L, Anderson D J, Engel Jr J and Buzsáki G 2000 Multiple site silicon-based probes for chronic recordings in freely moving rats: implantation, recording and histological verification Journal of neuroscience methods 98 77-82 [26] Fekete Z, Pongrácz A, Márton G and Fürjes P 2013b On the Fabrication Parameters of Buried Microchannels Integrated in In-Plane Silicon Microprobes Materials Science Forum 729 210-5 [27] Pongrácz A, Fekete Z, Márton G, Bérces Z, Ulbert I and Fürjes P 2013 Deep-brain silicon multielectrodes for simultaneous in vivo neural recording and drug delivery Sensors and Actuators B: Chemical [28] Frey O, Wal P D v d, Spieth S, Brett O, Seidl K, Paul O, Ruther P, Zengerle R and Rooij N F d 2011 Biosensor microprobes with integrated microfluidic channels for bi- directional neurochemical interaction Journal of Neural Engineering 8 066001 [29] Wu F, Stark E, Im M, Cho I-J, Yoon E-S, Buzsáki G, Wise K D and Yoon E 2013 An implantable neural probe with monolithically integrated dielectric waveguide and recording electrodes for optogenetics applications Journal of Neural Engineering 10 056012 [30] Kozai T D, Catt K, Li X, Gugel Z V, Olafsson V T, Vazquez A L and Cui X T 2015 Mechanical failure modes of chronically implanted planar silicon-based neural probes for laminar recording Biomaterials 37 25-39 [31] Mena-Segovia J, Sims H M, Magill P J and Bolam J P 2008 Cholinergic brainstem neurons modulate cortical gamma activity during slow oscillations The Journal of physiology 586 2947-60 [32] Márton G, Fekete Z, Fiath R, Baracskay P, Ulbert I, Juhász G, Battistig G and Pongrácz A 2013 In vivo measurements with robust silicon-based multielectrode arrays with extreme shaft lengths IEEE Sensors in press [33] Hajj-Hassan M, Chodavarapu V P and Musallam S 2009 Microfabrication of ultra-long reinforced silicon neural electrodes Micro & Nano Letters, IET 4 53-8 [34] Fekete Z, Hajnal Z, Márton G, Fürjes P and Pongrácz A 2013 Fracture analysis of silicon microprobes designed for deep-brain stimulation Microelectronic Engineering 103 160-6 [35] Moruzzi G and Magoun H W 1949 Brain stem reticular formation and activation of the EEG Electroencephalogr Clin Neurophysiol 1 455-73 [36] Camacho Evangelista A and Reinoso Suarez F 1964 Activating and Synchronizing Centers in Cat Brain: Electroencephalograms after Lesions Science 146 268-70 [37] Villanueva L, Desbois C, Le Bars D and Bernard J F 1998 Organization of diencephalic projections from the medullary subnucleus reticularis dorsalis and the adjacent cuneate 18 A robust silicon-based microelectrode array, mounted on a microdrive nucleus: a retrograde and anterograde tracer study in the rat The Journal of comparative neurology 390 133-60 [38] Winn P 2006 How best to consider the structure and function of the pedunculopontine tegmental nucleus: evidence from animal studies J Neurol Sci 248 234-50 [39] Rodrigo-Angulo M L, Rodriguez-Veiga E and Reinoso-Suarez F 2005 A quantitative study of the brainstem cholinergic projections to the ventral part of the oral pontine reticular nucleus (REM sleep induction site) in the cat Exp Brain Res 160 334-43 [40] Nuñez A, de Andres I and Garcia-Austt E 1991 Relationships of nucleus reticularis pontis oralis neuronal discharge with sensory and carbachol evoked hippocampal theta rhythm Exp Brain Res 87 303-8 [41] Nuñez A, Moreno-Balandran M E, Rodrigo-Angulo M L, Garzon M and De Andres I 2006 Relationship between the perifornical hypothalamic area and oral pontine reticular nucleus in the rat. Possible implication of the hypocretinergic projection in the control of rapid eye movement sleep Eur J Neurosci 24 2834-42 [42] Roš H, Magill P J, Moss J, Bolam J P and Mena-Segovia J 2010 Distinct types of non- cholinergic pedunculopontine neurons are differentially modulated during global brain states Neuroscience 170 78-91 [43] Vertes R P and Miller N E 1976 Brain stem neurons that fire selectively to a conditioned stimulus for shock Brain Research 103 229-42 [44] Steriade M, Dossi R C, Pare D and Oakson G 1991 Fast oscillations (20-40 Hz) in thalamocortical systems and their potentiation by mesopontine cholinergic nuclei in the cat Proceedings of the National Academy of Sciences of the United States of America 88 4396-400 [45] McCarley R W and Hobson J A 1971 Single neuron activity in cat gigantocellular tegmental field: selectivity of discharge in desynchronized sleep Science 174 1250-2 [46] Siegel J M, Wheeler R L and McGinty D J 1979 Activity of medullary reticular formation neurons in the unrestrained cat during waking and sleep Brain Research 179 49-60 [47] Tseng W T, Yen C T and Tsai M L 2011 A bundled microwire array for long-term chronic single-unit recording in deep brain regions of behaving rats Journal of neuroscience methods 201 368-76 [48] Paxinos G and Watson C 1998 The rat brain in stereotaxic coordinates: Academic Press) [49] Harris K D, Henze D A, Csicsvari J, Hirase H and Buzsaki G 2000 Accuracy of tetrode spike separation as determined by simultaneous intracellular and extracellular measurements Journal of neurophysiology 84 401-14 [50] Nicolelis M A L, Dimitrov D, Carmena J M, Crist R, Lehew G, Kralik J D and Wise S P 2003 Chronic, multisite, multielectrode recordings in macaque monkeys Proceedings of the National Academy of Sciences 100 11041-6 [51] Czurkó A, Hirase H, Csicsvári J and Buzsáki G 1999 Sustained activation of hippocampal pyramidal cells by 'space clamping' in a running wheel Eur J Neurosci 11 344-52 [52] Lőrincz M, Oláh M, Baracskay P, Szilágyi N and Juhász G 2007 Propagation of spike and wave activity to the medial prefrontal cortex and dorsal raphe nucleus of WAG/Rij rats Physiol Behav 90 318-24 [53] Gallyas F, Guldner F H, Zoltay G and Wolff J R 1990 Golgi-like demonstration of "dark" neurons with an argyrophil III method for experimental neuropathology Acta Neuropathol 79 620-8 19 A robust silicon-based microelectrode array, mounted on a microdrive [54] Gallyas F, Hsu M and Buzsáki G 1993 Four modified silver methods for thick sections of formaldehyde-fixed mammalian central nervous tissue: 'dark' neurons, perikarya of all neurons, microglial cells and capillaries Journal of neuroscience methods 50 159- 64 [55] Schuettler M 2007 Electrochemical Properties of Platinum Electrodes in Vitro: Comparison of Six Different Surface Qualities. In: Engineering in Medicine and Biology Society, 2007. EMBS 2007. 29th Annual International Conference of the IEEE, pp 186-9 [56] Toth A, Petyko Z, Mathe K, Szabo I and Czurko A 2007 Improved version of the printed circuit board (PCB) modular multi-channel microdrive for extracellular electrophysiological recordings J Neurosci Methods 159 51-6 [57] Otchy T M and Ӧlveczky B P 2012 Design and Assembly of an Ultra-light Motorized Microdrive for Chronic Neural Recordings in Small Animals Journal of Visualized Experiments : JoVE 4314 [58] Santos L, Opris I, Fuqua J, Hampson R E and Deadwyler S A 2012 A novel tetrode microdrive for simultaneous multi-neuron recording from different regions of primate brain Journal of neuroscience methods 205 368-74 [59] Muthuswamy J, Okandan M, Gilletti A, Baker M S and Jain T 2005 An array of microactuated microelectrodes for monitoring single-neuronal activity in rodents IEEE Trans Biomed Eng 52 1470-7 [60] Ward M P, Rajdev P, Ellison C and Irazoqui P P 2009 Toward a comparison of microelectrodes for acute and chronic recordings Brain Res 28 183-200 [61] Seymour J P and Kipke D R 2007 Neural probe design for reduced tissue encapsulation in CNS Biomaterials 28 3594-607 20
1704.02741
1
1704
2017-04-10T07:38:28
Avalanches and Generalized Memory Associativity in a Network Model for Conscious and Unconscious Mental Functioning
[ "q-bio.NC", "cond-mat.stat-mech" ]
We explore statistical characteristics of avalanches associated with the dynamics of a complex-network model, where two modules corresponding to sensorial and symbolic memories interact, representing unconscious and conscious mental processes. The model illustrates Freud's ideas regarding the neuroses and that consciousness is related with symbolic and linguistic memory activity in the brain. It incorporates the Stariolo-Tsallis generalization of the Boltzmann Machine in order to model memory retrieval and associativity. In the present work, we define and measure avalanche size distributions during memory retrieval, in order to gain insight regarding basic aspects of the functioning of these complex networks. The avalanche sizes defined for our model should be related to the time consumed and also to the size of the neuronal region which is activated, during memory retrieval. This allows the qualitative comparison of the behaviour of the distribution of cluster sizes, obtained during fMRI measurements of the propagation of signals in the brain, with the distribution of avalanche sizes obtained in our simulation experiments. This comparison corroborates the indication that the Nonextensive Statistical Mechanics formalism may indeed be more well suited to model the complex networks which constitute brain and mental structure.
q-bio.NC
q-bio
Avalanches and Generalized Memory Associativity in a Network Model for Conscious and Unconscious Mental Functioning Maheen Siddiqui∗ Centre for Brain and Cognitive Development, Birkbeck College, University of London, London WC1E 7HX, United Kingdom. Roseli S. Wedemann, Corresponding Author† Instituto de Matem´atica e Estat´ıstica, Universidade do Estado do Rio de Janeiro, Rua Sao Francisco Xavier 524, 20550-013, Rio de Janeiro, Brazil. Henrik Jensen‡ Centre for Complexity Science and Department of Mathematics, Imperial College London, London SW7 2AZ, United Kingdom, and Institute of Innovative Research, Tokyo Institute of Technology, 4259, Nagatsuta-cho, Yokohama 226-8502, Japan. (Dated: July 25, 2018) 7 1 0 2 r p A 0 1 ] . C N o i b - q [ 1 v 1 4 7 2 0 . 4 0 7 1 : v i X r a 1 Abstract We explore statistical characteristics of avalanches associated with the dynamics of a complex- network model, where two modules corresponding to sensorial and symbolic memories interact, representing unconscious and conscious mental processes. The model illustrates Freud's ideas regarding the neuroses and that consciousness is related with symbolic and linguistic memory activity in the brain. It incorporates the Stariolo-Tsallis generalization of the Boltzmann Machine in order to model memory retrieval and associativity. In the present work, we define and measure avalanche size distributions during memory retrieval, in order to gain insight regarding basic aspects of the functioning of these complex networks. The avalanche sizes defined for our model should be related to the time consumed and also to the size of the neuronal region which is activated, during memory retrieval. This allows the qualitative comparison of the behaviour of the distribution of cluster sizes, obtained during fMRI measurements of the propagation of signals in the brain, with the distribution of avalanche sizes obtained in our simulation experiments. This comparison corroborates the indication that the Nonextensive Statistical Mechanics formalism may indeed be more well suited to model the complex networks which constitute brain and mental structure. PACS numbers: Keywords: Consciousness-Unconsciousness, Neuroses, Self-organized Neural Networks, Boltzmann Machine, Generalized Simulated Annealing, Avalanches ∗Electronic address: [email protected] †Electronic address: [email protected] ‡Electronic address: [email protected] 2 I. INTRODUCTION The purpose of the present work is to define, measure and study statistical properties of avalanches which occur during memory retrieval in an artificial neural network, devel- oped to model neurotic phenomena and the associated conscious/unconscious interactions in mental life [1]. It was famously reported by Freud [2, 3] that patients with neurotic symptoms systematically repeated these symptoms in the guise of ideas and impulses. This tendency was referred to as the compulsion to repeat. The patients' repetitive behaviour suggested to Freud that they had ". . . the intention of correcting a distressing portion of the past. . . " [4]. Although the patient is aware of the obsessional ideas and impulses, and of the performance of the neurotic actions, the concomitant psychical predeterminants remain unconscious. In order to infer these predeterminants, and to bring them under the light of consciousness, Freud developed the analytical treatment called working-through. This treatment provides, through the interpretation of the predeterminants by freely talking in psychoanalytical sessions, the connections into which they are inserted. One of the early seminal findings of psychoanalytic research is that neurotic symptoms arise from traumatic and repressed memories. In fact, the repressed material can be con- strued as knowledge that, in spite of being present in the subject, is inaccessible to him through symbolical representation. In other words, this knowledge is momentarily or per- manently inaccessible to the patient's conscience, and is therefore regarded as unconscious knowledge [4, 5]. In these considerations, by symbolic representation we mean the association of symbols to meaning, as occurs in language, as well as in other ways of expressing thoughts and emotions, such as artistic representations (e.g., a painting or musical composition), and the recollection of dreams. Consequently, an important part of the analytical treatment proposed by Freud consists of a procedure for "bringing what is unconscious into consciousness". This is tantamount to a basic technique aiming at "filling up the gaps in the patient's memories, to remove his amnesias" [4]. Indeed, these amnesias are closely related to the origin of the neurotic symptoms, that is, to the compulsion to repeat. An important purpose of the therapeu- tic method called working-through, whereby neurotic analysands obtain relief and cure of painful symptoms, is to develop knowledge regarding the causes of those symptoms. This is achieved by accessing unconscious memories through free associative talking, which yields 3 understanding and a change in the analysand's compulsion to repeat [2, 3, 5]. The method is mainly based on the analysis of free associative talking, symptoms, parapraxes (slips of the tongue and pen, misreading, forgetting, etc.), dreams, and also on analyzing what is acted out in transference. This procedure allows the patient to slowly symbolize his repressed memories and to create new representations of his past experiences. An illustrative, schematic model of this neurotic mechanism and the working-through therapeutic method, was advanced in [1, 6]. It was there proposed that the neuroses can be understood in terms of an associative memory process in neural networks, where the network, when presented with an input pattern, retrieves a stored pattern which is most similar to the one currently shown. In order to model the compulsion to repeat a neurotic symptom, it was assumed that such a symptom is produced when the subject receives a stimulus resembling a repressed or traumatic memory trace. The stimulus then contributes to stabilize the neural net in a minimal energy state, corresponding to the trace that synthesizes the original repressed experience, which in turn produces a neurotic response (an act). The neurotic act does not result from the stimulus as a new situation, but as a response to the repressed memory. The model is based on the conception that the linguistic, symbolic, and associative pro- cess associated with psychoanalytic working-through therapy is mapped onto a correspond- ing scheme of reinforcing synapses among memory traces in the brain. These connections should involve declarative memory, implying that repressed memories are, at least par- tially, transformed into conscious ones. This takes into account the paramount importance that language has in psychoanalytic sessions, and the idea that unconscious memories are precisely those that cannot be expressed symbolically. It was thus proposed that, as the analysand symbolically elaborates manifestations of unconscious material through transfer- ence in psychoanalytic sessions, he creates new neural connections, and reinforces or inhibits older ones, reconfiguring the topology of his neural net. The network topology arising from this reconfiguration process stabilizes onto new energy minima, associated with new acts. Following the memory organization advanced in [7] (see also [1]), it is assumed that neu- rons belong to two hierarchically structured modules, respectively corresponding to sensorial and symbolic memories. Mental images of stimuli received by sensory receptors, either from the environment or from the body itself, are represented by memory traces stored in the sen- sorial memory module. On the other hand, higher level representations of traces in sensorial 4 memory, i.e. symbols, are stored in the symbolic memory module. This module represents brain structures corresponding to symbolic processing, language, and consciousness. Sen- sorial and symbolic memories are not isolated from each other and indeed, they interact, generating unconscious and conscious mental activity. In the model, the unconscious compulsion to repeat in neuroses [2 -- 4] is interpreted as a bodily response (an act) to an input stimulus (of any kind) that resonates with a pattern in sensorial memory, without activating symbolic memory. In this sense, the compulsion to repeat is akin to a reflexive act. This accounts for neurotic patients' claim that they cannot explain their neurotic acts. A sensorial memory trace becomes conscious when its retrieval also activates the retrieval of patterns in symbolic memory. If this happens, the resulting output does not resemble reflexive behaviour, and another level of processing becomes rele- vant. One may also have symbolic representations of emotions, as when someone says "I felt a warm happiness when I embraced my young nephew". Sensorial information that is not (or cannot be) associated to a symbol stays unconscious. The neurotic mechanism described here is in line with the hypothesis that the emergence of conscious experience requires the existence of a physical (neural) layer that is able to support metarepresentations [8]. A detailed description of the neural network model for these neurotic features can be found in [1], which includes a full description of the relevant algorithms. Memory functioning was first modelled by recourse to a Boltzmann Machine (BM), which is a stochastic extension of the Hopfield model. However, it was later found that the node-degree distributions of the hierarchically clustered network topologies generated by the model's clustering algorithm (see section II), with long and short-range synapses, are well represented by the asymp- totic power-law, q-exponential distributions. This behaviour indicates that the statistical features of our model may not be well described by Boltzmann-Gibbs statistical mechan- ics, but rather by Nonextensive Statistical Mechanics (NSM) [9 -- 11]. The NSM theoretical framework, and its variegated applications to physics, biology, economics, and other areas, have been the focus of an intense research activity in recent years [10]. The concomitant formalism has been applied to the study of diverse types of complex systems, including sys- tems with long-range interactions [12], systems exhibiting weak chaos [13], complex networks [14], processes involving nonlinear diffusion [15], and many others. In recent work [16, 17] (see also references therein), the authors have also used Hopfield neural networks to model attachment in developmental psychology, as well as behavioural patterns in psychology and 5 psychotherapy. In the present work, we thus use the model described in [1], where memory is simulated by a generalization of the BM inspired on NSM, called Generalized Simulated Annealing (GSA) [1, 9], and this affects the sequence of associations of thoughts in the mental processes we are illustrating. The model is in good qualitative agreement with the main facts provided by psycho- analytic experience. In particular, it is consistent with the (sometimes exasperating) slow nature of the working-through process. The model's dynamics reproduces in a plausible way the re-association of unconscious sensorial memory traces, and of new experiences, to symbolic processing areas, mimicking the repetitive, adaptive, reinforcement learning in- volved in the simulation of working-through. As a result of this self-reconfiguration process, represented in the model by a change in network connectivity, the analysand is partially freed from his/her original neurotic states and concomitant acts. The new network topology evolves to, and stabilizes itself onto, new energy minima. These resulting network states are associated with new conscious or unconscious acts. Evidently, the ultimate aim of both the analysand and the analyst is to reach new states, and generate new acts, that are more pleasant and comfortable to the analysand and his relations. As the therapy progresses, the analysand rewrites his/her life history, through a process of new significations. Furthermore, and perhaps more importantly, both the present and the future of the analysand are also being rewritten, by creating new possibilities, and by opening new windows of opportunity for the pursuit of happiness. Our central tenet is that this story is embedded, i.e. written, in the individual's biological neural network. It is clear that psychoanalytic theory is still far from having the rigorous quantitative support required by modern science. However, we agree with some contemporary scien- tists [18, 19], as well as early psychoanalysts [4, 20, 21] that, although this state of affairs poses serious limitations and difficulties, it also constitutes a challenge and a stimulus for further scientific research. These venues of enquiry are worth pursuing, since many find- ings of psychodynamic theories have already contributed both to our understanding and characterization of mental phenomena, as well as to the development of successful clinical treatments for many mental disorders [18, 19]. In the present work, we study some additional properties of the aforementioned model, which give us further knowledge on how basic microscopic and macroscopic features and mechanisms influence emergent behaviour of the complex network structures proposed by 6 the model. In Section II, we give a brief overview of the basic algorithms that characterize the model and which we use in our simulation experiments. We present a definition of avalanche sizes during the memory retrieval mechanism and results of their measurements, in simulation experiments in Section III. These avalanche sizes can be qualitatively compared to brain imaging experiments [22], showing that the NSM memory retrieval mechanism produces power-law behaviour, which does not emerge from BM functioning, i.e. from Boltzmann-Gibbs statistical mechanics. II. NETWORK TOPOLOGY AND MEMORY ACCESS MECHANISMS The topological structure of each of the two memory modules was generated by a clus- tering algorithm proposed in [7] (see also [1]), which models the self-organizing process that controls synaptic plasticity, resulting in a hierarchically structured neural network topol- ogy. The algorithm is inspired on microscopic biological mechanisms, found in typical brain, cellular processes of many animals [23, 24]. As an example we can mention the on-center/off-surround structure, characterized by neurons that are in cooperation, through excitatory synapses, with other neurons in their immediate neighbourhood, while they are in competition with neurons that lie outside these surroundings. Competition and cooperation occur both in statically hardwired structures, and as part of a variety of neuronal dynamical processes, where neurons compete for certain chemicals [23, 24]. In synaptogenesis, for instance, stimulated neurons release neural growth factors that spread through diffusion, reaching neighbouring cells and promoting synaptic growth. The cells receiving neural growth factors make synapses and live, while those hav- ing no contact with these substances die [23, 25]. A neuron releasing neural growth factors ushers the process of synaptic formation in its tridimensional neighbourhood, and becomes a center of synaptic convergence. Neighbouring neurons releasing different neural growth factors at different rates give rise to several synaptic convergence centers. These centers then compete through the new synapses being created in their surroundings. Through these processes, a signaling network is established, that controls the development and plasticity of the brain's neuronal circuits. This neural competition is started and controlled by envi- ronmental stimulation and, consequently, it constitutes an important mechanism through which features of the environment can be mapped onto brain structures. 7 A. Clustering Algorithm We reproduce the algorithm proposed in [1, 7] here, to aid understanding of the mea- surements and analysis which we introduce in this paper. This clustering algorithm models the self-organizing process which controls synaptic plasticity, and results in a structured hierarchical topology of each of the two memory modules. It consists of the following steps. Step 1 The initial bidimensional positions of the neurons are randomly generated according to a uniform probability density on a square sheet. Step 2 To simulate synaptic growth, we assume a Gaussian solution of the equation gov- erning the diffusion of the neural growth factors, thus avoiding the time-consuming numerical treatment of this equation. Consequently, a synapse with strength wij (the synaptic weight) is allocated to transmit the output signal from neuron nj to a neuron ni, according to the Gaussian probability density Pij = exp(−(rj − ri)2/(2σ2))/√2πσ2 , (1) where rj and ri are the respective positions of nj and ni in the bidimensional sheet, and σ is the standard deviation of the Gaussian distribution, which is here a model parameter. If a synapse connecting nj to ni is generated, its strength wij is proportional to Pij. Step 3 It was verified in [26] that cortical maps representing different stimuli are formed, such that each stimulus activates a group of neurons spatially close to each other, and that these groups are uniformly distributed along the sheet of neurons representing memory. So one now randomly selects m neurons which will each be a center of the representation of a stimulus. To choose the value of m, one should take into account the storage capacity of the BM [27]. Step 4 Reinforce synapses adjacent to each of the m centers chosen in Step 3, according If ni is a center, define sumni = Pj wij. For each nj to the following criteria. adjacent to ni, increase wij by ∆wij, with probability P robnj = wij/sumni, where ∆wij = ηP robnj and η ∈ ℜ is a model parameter chosen in [0, 1]. After incrementing wij, decrement ∆wij from the weights of all the other neighbours of ni, according to: ∀k 6= j,wik = wik − ∆wik, where ∆wik = (1 − wik/Pk6=j wik)∆wij. 8 Step 5 Repeat step 4 until a clustering criterion is met. In the above clustering algorithm, Step 4 regulates the strength of synaptic connections, i.e., plasticity, by intensifying synapses within a cluster and reducing synaptic strength between clusters (disconnecting clusters). A cluster is therefore formed by a group of neu- rons that are close to each other, with higher probability of being connected by stronger synapses. This mechanism is akin to a preferential attachment criterion, constrained by an energy conservation (neurosubstances) prescription, controlling synaptic plasticity. Neurons that have been sensorially more stimulated and are therefore more strongly connected will stimulate other neurons in their neighbourhoods and promote still stronger connections. This is compatible with the biological mechanisms mentioned earlier. The growth of long-range synapses in the brain occurs less frequently than the growth of short-range ones. The reason for this is that the former are energetically more costly than the latter. In order to allocate long-range synapses connecting clusters, one should regard the basic learning scheme proposed by Hebb [27 -- 29], which is based on the idea that synaptic growth among two neurons is promoted by their concomitant stimulation. Since we are still not aware of the synaptic distributions that result in topologies which represent the structure of associations of symbols in language and thought, as a first approximation, we have allocated long-range synapses randomly among clusters of neurons (for a more detailed discussion see [1, 7]). Within a randomly chosen cluster C, defined by one of the m neurons which is the center of representation of a stimulus (step 3 of the clustering algorithm), a neuron ni is chosen to receive a long-range connection with probability Pi = Pj wij/Pnj ∈CPk wjk. If the long-range synapse connects clusters in different memory sheets (sensorial and symbolic memories), its randomly chosen weight is multiplied by a real number ζ in the interval (0, 1], reflecting the fact that, in neurotic patterns, sensorial information is weakly accessible to consciousness, i.e., repressed. B. Memory Retrieval The topologies generated with the clustering algorithm have a hierarchical structure and the average node-degree distributions present an asymptotic power-law behaviour [1]. The functioning of memory retrieval was originally modelled by a Boltzmann Machine (BM) [27, 30]. There is no theoretical indication of the exact relation between network 9 topology and memory access dynamics. There have been indications that complex physical systems characterized by spatial disorder and/or long-range interactions, often presenting power-law behaviour (are asymptotically scale invariant) may be described by the Nonex- tensive Statistical Mechanics (NSM) formalism [9 -- 15]. The power-law and generalized q- exponential behaviour for the node-degree distributions of the network topologies generated be the clustering algorithm [1] indicate that they may not be well described by Boltzmann- Gibbs (BG) statistical mechanics, but rather by NSM [9]. Memory access was thus modelled by a generalization of the BM called Generalized Simulated Annealing (GSA) [1, 9], and this changes the chain of associations generated by the model. In the BM [27, 30], the N nodes in the neural network are connected symmetrically by weights wij = wji. The state Si of each unit ni takes output values in {0, 1}. As a consequence of the symmetry of the connections, one can associate an energy functional H({Si}) = − wijSiSj , 1 2Xij (2) to network state S = {Si} and, according to the BG distribution, the transition probability (acceptance probability) from state S to S′, if H(S′) ≥ H(S), is given by PBG(S → S′) = exp(cid:20) H(S) − H(S′) T (cid:21) , (3) PGSA(S → S′) =  where T is the network "temperature" parameter. Pattern retrieval on the net is achieved by a standard simulated annealing process, in which T is gradually lowered by a factor α, according to the BG distribution [27]. In the NSM formalism, one uses a generalized acceptance probability [9] for a transition from S to S′, if H(S′) ≥ H(S), given by 1 [1+(qA−1)(H(S ′)−H(S))/T ]1/(qA −1) , 0 , if ϕ > 0 , if ϕ ≤ 0 , (4) where qA is a parameter called q-acceptance and ϕ = 1 + (qA − 1)(H(S′) − H(S))/T . In If one the limit qA → 1, (4) reduces to the Boltzmann-Gibbs transition probability (3). uses transition probability (4) in place of transition probability (3), in the standard BM simulated annealing algorithm, the resulting procedure is called GSA (see [9] for a more detailed discussion). It is convenient that we define, for each transition from state S to S′ during annealing, the quantity ∆E = H(S′) − H(S) . 10 (5) Both the BM and GSA differ from a gradient descent minimization scheme since, besides allowing state transitions that lower the total energy of the network (2), they also allow the system to transition into a state with an increase in energy, depending on the values of T and qA, according to (3) and (4). The BG transition probability (3) predominantly allows changes of states with small increases in energy, and state transitions with higher energy increases occur with almost negligible probabilities. The BM will thus strongly prefer visiting state space within a nearby energy neighbourhood from the starting point (initial state of the annealing process). The GSA transition probability (4) allows state transitions with higher energy increases than the BM and, although the probabilities for these transitions are very low, they are still considerable when compared to the BM. This allows the system to transition into attractor states that are farther in state space from the initial network state and also into basins of attraction corresponding to higher energy values [1, 31] (see Figure 2 in [31]). This implies that the hierarchically structured memory modules having both long and short-range connections, with memory access modelled by GSA achieves associations among memory traces that are not achieved by the BM [31]. This increase in associativity observed in the GSA memory retrieval mechanism, when compared to the BM, suggests that if memory functions according to the NSM theoretical framework, one will have a more creative mode of memory and mental functioning. In traditional neural network modelling, the temperature parameter is inspired by the fact that biological neurons fire with variable strength, and that there are delays in synapses, random fluctuations from the release of neurotransmitters, and so on. These effects can be considered as noise in synaptic functioning [23, 27], and we may thus consider that temperature in the BM and GSA and the qA parameter in GSA control noise. In the model we are considering [1], non-zero temperature and qA 6= 1 values regulate associations among memory traces, in an analogy with the concept that freely talking in analytic sessions and stimulation from the psychoanalyst lower resistances and allow greater associativity and creativity. Once the network topology is generated by the clustering algorithm, one can find the stored patterns by presenting many random patterns to both the BM and the GSA mech- anism, with an annealing schedule α that allows stabilizing onto the many local minimum values of the network energy function. Each of the minimum energy values corresponds to a stable state of the network and is associated with a stored memory trace. These initially 11 stored patterns represent the neurotic memory attractors (as in the compulsion to repeat), since they are associated with the two weakly linked sensorial and symbolic subnetworks. In [26], Carvalho et al. proposed a neurocomputational model to describe how the original memory traces are formed in cortical maps. It is important to note that in our experiments, we are not aiming at finding a global minimum energy state, but at visiting the many local minima of the energy landscape, which represent stored information in a cognitive network. III. AVALANCHES IN MEMORY ACCESS In the present work, we have defined and measured avalanche sizes during the memory retrieval process. The size of an avalanche during the simulated annealing procedure, both in the BM and in GSA, is defined to be the number of state changes that occur during annealing, from the initial to the final network state, which corresponds to one of the minima in the energy landscape of the network. This is equivalent to saying that the size of an avalanche corresponds to the number of hops (steps) that the access mechanism performs on the network energy landscape, from an initial state until it reaches one of the local minima. The avalanche size should thus be proportional to the time for the propagation of a signal (stimulus), during access of information in memory. In other words, the avalanche size should be related to the time associated with memory retrieval and also to the size of the neuronal region activated during memory access. In each of the simulation experiments we present here, we performed 2,048,000 minimiza- tion (annealing) procedures, starting each one from a different random network configura- tion. Since the simulations require much computational time (many days and even a few weeks), we have analysed smaller networks with total number of neurons N = 32, such that Nsens = Nsymb = 16 neurons belong to the sensorial and symbolic modules, respectively. This is a small network size when one considers the brain, as there are billions of neurons in the human brain. However, the brain is considered to be scale invariant [22] and because the short-range mechanisms in the algorithms we use to create network topology are scal- able, we expect that our experiments should, to some extent, qualitatively be comparable to biological processes [1]. The annealing schedule was controlled so that the network stabilizes on the many local minima of the energy landscape, and was the same for both machines in all experiments. We 12 conducted experiments with different values of the temperature T and qA parameters. Both of these parameters model stochastic fluctuations in network functioning and we studied how the change in T and qA values caused a change in the behaviour of the two machines, thus affecting the distribution of avalanche sizes, the network energy loss and measurements of correlations among energy increments along the path followed during annealing. In Figure 1, we show the avalanche size distributions obtained with the two machines, for T = 0.05 and qA = 1.3. In this experiment, when executing step 2 of the clustering algorithm, 50% of the synapses of the network are excitatory (positive synaptic weights) and 50% are inhibitory (negative synaptic weights). On the left, we see the avalanche size distributions for the BM and on the right, for GSA. From the distribution in Figure 1-a, it is clear that the BM has a preferred avalanche size of 125 with an associated probability of occurrence of 0.96. This distribution also has a smaller peak at avalanche size 1325, which occurs with probability 0.0022. The smaller peak has a slower decay before a sharp drop. The distribution for GSA in Figure 1-b, on the other hand, behaves quite differently. For the range of avalanche sizes we measured, we observe a monotonic drop in the fre- quency for GSA. The points measured in this simulation can be approximately fitted by a q-exponential [10] expq(x) = [1 + (1 − q)x] 0 ,  1 1−q , if 1 + (1 − q)x > 0 , if 1 + (1 − q)x ≤ 0 . (6) This function is at the core of nonextensive thermostatistics, being the result of the con- strained optimization of the power-law, nonadditive entropic functional Sq = 1 q − 1 [1 −Xi pq i ] , (7) where pi is a normalized probability distribution and q is a real parameter. The q-exponential asymptotically behaves as a power-law and, in Figure 1-b, the curve which fits the points within the observed values of avalanche sizes, corresponding to q = qf = 1.19, does show an asymptotic behaviour akin to a power-law. The imaging technique in neuroscience referred to as functional magnetic resonance imag- ing (fMRI) uses a method known as blood-oxygen-level dependent (BOLD) contrast imag- ing [32]. Most cells in the body possess their own reservoirs of sugar which undergoes respiration to produce energy, when the cells need it. It is known that neurons do not 13 101 100 10-1 10-2 10-3 10-4 10-5 10-6 y c n e u q e r F 10-7 10 BM, T = 0.05 100 1000 10000 Avalanche Size (a) 102 101 100 10-1 10-2 10-3 10-4 10-5 10-6 10-7 y c n e u q e r F 10-8 100 GSA, T = 0.05, qA = 1.3 0.09*(( 1-(1-1.19)*( (x - 360) /112.0 ) )**(1/(1-1.19))) 1000 Avalanche Size (b) 10000 FIG. 1: (a) Frequency of occurrence of avalanche sizes for the BM. (b) Frequency of avalanche sizes with GSA. For both machines, T = 0.05 and for GSA, qA = 1.3. In both cases, 50% of the synapses are inhibitive. All depicted quantities are dimensionless. possess internal stores of energy and, when a neuron is in a firing state it requires energy in the form of sugar and oxygen, which it obtains by means of a haemodynamic response, whereby blood releases oxygen to the firing neurons at a greater rate in comparison to those in resting state. This causes a difference in magnetic susceptibility between oxygenated and deoxygenated blood, which results in a magnetic signal variation that may be detected on an MRI scanner [32]. Studying changes in the brain BOLD signal allows observation of different active areas of the brain to further understand certain aspects of brain functioning. In 2003, the work of Beggs and Plenz [33] with cortical slices revealed the phenomena which they called neuronal avalanches in the brain. In those observations, avalanches were defined as short bursts of activity that last a few milliseconds, followed by several seconds of inactivity. These phenomena can be observed in cortical slices of the neocortex, although their relation to physiological processes in the brain is still unknown [33]. The avalanche sizes detected by Beggs and Plenz follow a power-law-like distribution, with an exponent of -3/2. In their observations, they recorded spontaneous local field potentials continuously using a 60 channel multielectrode array, and avalanches are a measure of time duration and spatial reach of the propagation of signals in the brain. Tagliazucchi et al. [22] extended the work done by Beggs and Plenz [33], "by inspecting only the relatively large amplitude BOLD signal peaks, suggesting that relevant information 14 can be condensed in discrete events". The authors present a spatiotemporal point process, where the timing and location of these discrete events is obtained to study and capture the dynamics of the brain in a resting state. When regions of the brain show activity above a certain threshold level, the detected BOLD signal determines the location of the activity as well as the duration. Figure 3-A in [22] shows examples of co-activated clusters, defined as "... groups of contiguous voxels with signal above the threshold at a given time" (clusters are measured in units of voxels). In [22], Figure 3-D shows the distribution of average measured cluster sizes in fMRI im- ages for ten individuals, which has a power-law (scale-free) behaviour spanning four orders of magnitude, also with an exponent of approximately -3/2. The size of a cluster is propor- tional to the duration of brain activity. The longer the duration for which an avalanche is seen in the brain, the larger are the observed cluster sizes. This happens because when the duration of activity is longer, clusters (voxels) cause the activation of neighbouring clusters, before reducing activity and fading away. This behaviour is comparable to the avalanche size distributions produced by our computational model, as in Figure 1-b, because an avalanche size, in our simulation experiments, was defined in a way so that it should be proportional to the time for propagation of a signal during the access of information in memory. This is a reasonable assumption, since transition probabilities (3) and (4) which regulate network functioning should mimic the way the memory network occupies phase-space, during the re- trieval of a mnemenic trace. So the larger avalanche sizes observed in our simulations should be associated with larger cluster sizes, as a larger avalanche size is associated with a longer time for propagation of a signal, with access to a larger number of neighbouring neurons. We therefore qualitatively compared the behaviour of the distribution of cluster sizes obtained in [22] with the distribution of avalanche sizes obtained in our simulation experiments, for some values of the T and qA parameters, and we found q-exponentials with a similar asymp- totic power-law like behaviour, spanning similar orders of magnitude of avalanche sizes and frequencies, only for the GSA machine, however with larger absolute values of the exponent (exponent close to -5 in Figure 1-b and approximately -10 in Figure 4-b). The Boltzmann machine did not present power-laws in our experiments. This qualitatively corroborates our initial indication, that the power-law and generalized q-exponential behaviour of quantities which describe the structure of the complex networks generated by the clustering algorithm of our model suggest that they may not be well described by Boltzmann-Gibbs (BG) statis- 15 tical mechanics, but rather by NSM [1, 9]. For a sequence of values of energy increments ∆E given by (5) during an avalanche, we also measured if the value ∆E(τ0) at step τ0 influences (is correlated to) the value at a later step ∆E(τ0 + τ ). The temporal correlation function G(τ ) among elements in a sequence separated by an interval τ is defined as [34] G(τ ) = m−τ Xτ0=1 ∆E(τ0)∆E(τ0 + τ ) m − τ − m Xτ0=1 ∆E(τ0) m !2 , (8) where m is the size of the sequence which is being considered, and should be much larger than the largest value of τ in the domain of (8). In Figure 2, we show correlations for one avalanche of the BM in Figure 2-a and for another avalanche generated by GSA in Figure 2-b. Since avalanches in the BM and GSA typically have different sizes, Figures 2-a and b correspond to different values of m. For both machines we notice an approximate exponential fall of G(τ ) for smaller values of τ , with larger fluctuations around an average decay for GSA. These larger fluctuations for GSA occur because GSA generates higher probabilities of making state transitions with positive values of ∆E than the BM. After a rapid decrease, both figures converge to a situation where they fluctuate around zero correlation values. A similar behaviour occurs for the other avalanches which we measured. For larger values of T and qA, the fluctuations around average correlation values are much larger, since the machines then have higher probabilities of making state transitions with positive values of ∆E, and the first term in (8) is negative more often than for lower values of T and qA. We show avalanche size distributions obtained with the two machines, for T = 0.2 and qA = 1.6, in Figure 3. As a consequence of the increase of the values of the two parameters that control associativity and noise in the network, both machines have a higher probability of increasing energy during the annealing process, and therefore achieve higher values of avalanche sizes more frequently. The BM now has a preferred avalanche size of 1025, with a corresponding probability of 0.31, as seen from Figure 3-a. The distribution of avalanche sizes for GSA now loses the general q-exponential behaviour and has two preferred values (the first is 4775 with 0.035 probability and the second is 6375 with probability 0.056), with a very short power-law like behaviour after the second peak, before a sharp drop. The increase in frequency of larger avalanche sizes for both machines is consistent with the increase in associativity for larger values of T and qA found in [31], with GSA still producing more 16 ) τ ( G n o i t l a e r r o C 0.035 0.03 0.025 0.02 0.015 0.01 0.005 0 -0.005 0 GSA, Avalanche Size = 2873 BM, Avalanche Size = 1593 0.07 0.06 0.05 0.04 0.03 0.02 0.01 0 ) τ ( G n o i t l a e r r o C 5 10 τ (a) 15 20 25 -0.01 0 5 10 15 25 30 35 40 20 τ (b) FIG. 2: (a) Correlations among the sequence of ∆E values during an avalanche of the BM. The avalanche size for this experiment is m = 1593 and the maximum value of τ is 25. (b) Correlations among the ∆E for an avalanche generated by GSA. The avalanche size for this experiment is m = 2873 and the maximum value of τ is 40. For both machines, T = 0.05 and for GSA, qA = 1.3. 50% of the synapses are inhibitive. All depicted quantities are dimensionless. associativity than the BM. There has been recent work regarding the study of the proportion of inhibitory synapses in the brain (see [35] and references therein). In mammals, this proportion has been measured to have values between 20 and 30%. We have thus conducted avalanche measurements in our model so that, when executing step 2 of the clustering algorithm, 70% of randomly chosen synapses of the network are excitatory (positive synaptic weights) and 30% are inhibitory (negative synaptic weights). A distribution of avalanche sizes for this proportion of inhibitory synapses, for both the BM and GSA with T = 0.2, and qA = 0.7 in GSA is shown in Figure 4. Again we see a preferred avalanche size for the BM, in this case at 1025 with probability 0.32. We also observe once more for GSA in Figure 4-b, for the range of avalanche sizes which we measured, a monotonic decrease in the distribution of avalanche sizes, so that the experimental points resulting from this simulation can be approximately fitted by a q-exponential, with q = qf = 1.098. This is also asymptotically in accordance with the scale-free, power-law behaviour observed experimentally in [22]. Besides the avalanche size distributions and temporal correlations, we also measured the 17 100 10-1 10-2 10-3 10-4 10-5 10-6 y c n e u q e r F 10-7 10 GSA, T = 0.2, qA = 1.6 BM, T = 0.2 10-1 10-2 10-3 10-4 10-5 10-6 y c n e u q e r F 100 1000 10000 Avalanche Size 10-7 3000 4000 5000 6000 Avalanche Size 7000 8000 9000 10000 (a) (b) FIG. 3: (a) Frequency of occurrence of avalanche sizes for the BM. (b) Frequency of avalanche sizes with GSA. For both machines, T = 0.2 and for GSA, qA = 1.6. In both cases, 50% of the synapses are inhibitive. All depicted quantities are dimensionless. y c n e u q e r F 101 100 10-1 10-2 10-3 10-4 10-5 10-6 10-7 BM, T = 0.2 101 100 10-1 10-2 10-3 10-4 10-5 y c n e u q e r F 10-6 100 GSA, T = 0.2, qA = 0.7 0.35*(1-(1-1.098)*((x - 160)/161.0))**(1/(1-1.098)) 1000 Avalanche Size (b) 100 Avalanche Size 1000 (a) FIG. 4: (a) Frequency of occurrence of avalanche sizes for the BM. (b) Frequency of avalanche sizes with GSA. For both machines, T = 0.2 and for GSA, qA = 0.7. In both cases, 30% of the synapses are inhibitive. All depicted quantities are dimensionless. distributions of the total energy lost by the network, during the annealing processes of each set of simulation experiments, for both machines, i.e. H(SF inal) − H(SInit) , where SF inal and SInit are respectively the final and initial states of the network, during one memory retrieval procedure. These distributions are basically the same for both machines during the 18 same experiment, with just a very slight and negligible difference close to the tales of the distributions, due to limited size statistics. IV. CONCLUSIONS We have further studied the memory mechanism in a schematic neural network model, which illustrates conscious and unconscious mental processes in neurotic mental function- ing [1 -- 5]. The model illustrates aspects of mental functioning related to memory associa- tivity, creativity, consciousness and unconsciousness, which are present both in pathological and normal mental processing. The model is based on a self-organizing, clustering algorithm which generates a hierarchically structured bimodular neural network, based on basic biolog- ical cellular mechanisms, which have aspects akin to a preferential attachment mechanism, constrained by an energy conservation (conservation of neurosubstances) prescription, con- trolling synaptic plasticity. The networks generated by the algorithm have a small-world-like topology, with clusters of neurons strongly coupled by short-range connections (synapses) and also less frequent long-range connections. The node-degree distributions of these net- work topologies present a generalized q-exponential and asymptotic power-law behaviour. This structure suggests that memory access may best be modelled by a generalization of the Boltzmann Machine called Generalized Simulated Annealing, and this determines the chain of associations generated by the model. In the present work, we have introduced a definition of avalanche sizes during the simu- lated annealing procedure of the BM and GSA machine, which model memory access. Both the BM and GSA are regulated by parameters (temperature T and qA for the NSM formal- ism) which control stochastic fluctuations in network functioning and model noisy behaviour present in biological synaptic mechanisms. We conducted simulation experiments with dif- ferent values of T and qA, and studied how this causes a change in the behaviour of the two machines, thus affecting the distribution of avalanche sizes, the network energy loss and measurements of correlations among energy increments along the path followed during an- nealing. In these experiments, we found that for some values of T and qA, the avalanche size distributions of the GSA machine may be approximately fitted by a q-exponential, with the corresponding asymptotic power-law behaviour. The BM does not generate this behaviour and produces smaller avalanche sizes, related to less associative memory functioning, when 19 compared to GSA. We conjecture that avalanche sizes defined for our model should be related to the time consumed and also to the size of the neuronal region which is activated, during access of information in memory. This assumption is reasonable, since the transition probabilities (3) and (4) which regulate the BM and GSA network functioning should mimic the way the memory apparatus occupies phase-space, during the retrieval of a mnemenic trace. We may therefore qualitatively compare the behaviour of the distribution of cluster sizes, obtained during fMRI measurements of the propagation of signals in the brain reported in [22], with the distribution of avalanche sizes obtained in our simulation experiments. For some values of the T and qA parameters, we find a similar power-law like behaviour, spanning similar orders of magnitude of avalanche sizes and frequencies, only when using the GSA machine, which is based on the NSM formalism. The BM did not produce this pattern in our simulation experiments. This result experimentally corroborates our original indication, that the q- exponential and asymptotic power-law (scale-free) behaviour of macroscopic quantities which describe the structure of the complex networks generated by the algorithm of our model suggest that they may indeed be better described by the NSM [9] formalism, rather than by Boltzmann-Gibbs (BG) statistical mechanics. We have also considered recent indications that the brains of mammals are composed of a proportion of inhibitive synapses, which has been measured to have values between 20 and 30% [35]. We therefore measured avalanche size distributions during memory ac- cess, in networks generated by the clustering algorithm of our model with a proportion of 30% of inhibitive synapses. In this case, we also found that, for some T and qA values, the avalanche size distributions may be fitted by a q-exponential (with the corresponding asymptotic power-law behaviour), only for the GSA machine. Indeed, the q-exponential fit for the case we analysed seems even better with this proportion of inhibitive synapses than with 50% of inhibitive synapses. Although the fMRI measurements presented in [22], capture the dynamics of the brain in a resting state, the authors argue that "... despite the fact that in resting data there are not explicit inputs, the average BOLD signal around the extracted points ... still resembles the hemodynamic response function (HRF) evoked by a stimulus". This further supports the qualitative comparison of the results of our simulation experiments with their fMRI measurements. 20 We may thus conclude that the computational model studied in this work is viable as an illustrative schematic model of some basic mental mechanisms and in fact is qualitatively comparable to fMRI data obtained from patients, demonstrating actual brain activity. This comparison also reinforces the idea that complex systems such as the brain may well be better described by Nonextensive Statistical Mechanics, which produces the asymptotic power-law behaviour, for various temperatures and qA values, while Boltzmann-Gibbs sta- tistical mechanic does not show such behaviour. In [33], the authors argue that in the critical states characterized by power-laws, "the network may satisfy the competing demands of in- formation transmission and network stability". As demonstrated in earlier work regarding this model [1, 31], as well as in our simulations, memory retrieval governed by Nonex- tensive Statistical Mechanics allows much more associativity in memory functioning than Boltzmann-Gibbs statistical mechanics, while still achieving the stability necessary for infor- mation storage (see also [36]). This results in a potentially more creative mental structure with more capacity for the establishment of metaphors. From the perspective of psychother- apy, patients with more creativity have a larger possibility of reassociating traumatic and repressed experiences, and to construct new historical perspectives for their present and fu- ture, during therapy. Further investigations along this line of research are important, since they may reveal more basic structural features of brain and mental functioning. Acknowledgments We are grateful to Constantino Tsallis, Evaldo Curado, Angel R. Plastino and Abbas Edalat for fruitful conversations. This research was developed with grants from the Brazil- ian National Research Council (CNPq), the Rio de Janeiro State Research Foundation (FAPERJ) and the Brazilian agency which funds graduate studies (CAPES). [1] R.S. Wedemann, R. Donangelo, L.A.V. de Carvalho, Generalized Memory Associativity in a Network Model for the Neuroses, Chaos 19 (2009) 015116-(1 -- 11). [2] S. Freud, Beyond the Pleasure Principle, Standard Edition of The Complete Psychological Works of Sigmund Freud, Vol. XI, First German edition in 1920, The Hogarth Press, London, 1974. 21 [3] S. Freud, Remembering, Repeating and Working-Through, Standard Edition of The Complete Psychological Works of Sigmund Freud, Vol. XII, First German edition in 1914, The Hogarth Press, London, 1953. [4] S. Freud, Introductory Lectures on Psycho-Analysis, Standard Edition, First German edition in 1917, W. W. Norton and Company, New York - London, 1966. [5] S. Freud, The Unconscious, Standard Edition of The Complete Psychological Works of Sig- mund Freud, First German edition in 1915, Vol. XIV, The Hogarth Press, London, 1957. [6] R.S. Wedemann, R. Donangelo, L.A.V. de Carvalho, I.H. Martins, Memory Functioning in Psychopathology, Lecture Notes in Computer Science 2329 (2002) 236 -- 245. [7] R.S. Wedemann, L.A.V. de Carvalho, R. Donangelo, A Complex Neural Network Model for Memory Functioning in Psychopathology, Lecture Notes in Computer Science 4131 (2006) 543 -- 552. [8] A. Cleeremans, B. Timmermans, A. Pasquali, Consciousness and metarepresentation: A com- putational sketch, Neural Networks 20 (2007) 1032 -- 1039. [9] C. Tsallis, D.A. Stariolo, Generalized simulated annealing, Physica A 233 (1996) 395 -- 406. [10] C. Tsallis, Introduction to Nonextensive Statistical Mechanics, Approaching a Complex World, Springer, New York, 2009. [11] C. Beck, Generalised information and entropy measures in physics, Contemporary Physics 50(4) (2009) 495 -- 510. [12] C. Vignat, A. Plastino, A.R. Plastino, Entropic upper bound on gravitational binding energy, Physica A 390 (2011) 2491 -- 2496. [13] U. Tirnakli, E.P. Borges, The standard map: From Boltzmann-Gibbs statistics to Tsallis statistics. Nature Sci. Rep. 6 (2016) 23644.1 -- 8. [14] S. Brito, L.R. da Silva, C. Tsallis, Role of dimensionality in complex networks. Nature Sci. Rep. 6 (2016) 27992.1 -- 8. [15] M.S. Ribeiro, F.D. Nobre, E.M.F. Curado, Classes of N-Dimensional Nonlinear Fokker-Planck Equations Associated to Tsallis Entropy, Entropy 13(11), (2011) 1928 -- 1944. [16] A. Edalat, F. Mancinelli, Strong Attractors of Hopfield Neural Networks to Model Attachment Types and Behavioural Patterns, in: P. Angelov, D. Levine (Eds.), Proceedings of The 2013 International Joint Conference on Neural Networks (IJCNN), IEEE - Curran Associates, Inc., New York, 2013, pp. 1-10. 22 [17] A. Edalat, F. Mancinelli, Introduction to self-attachment and its neural basis, in: D.S. Huang, Y. Choe (Eds.), Proceedings of The 2015 International Joint Conference on Neural Networks (IJCNN), IEEE - Curran Associates, Inc., New York, 2015, pp. 1-8. [18] E. Kandel, Psychiatry, Psychoanalysis, and the New Biology of Mind, American Psychiatric Publishing, Inc., Washington D.C., London, 2005. [19] Shedler, J., The Efficacy of Psychodynamic Psychotherapy, American Psychologist 65(2) (2010) 98 -- 109. [20] Lacan, J.: On a Discourse that Might not be a Semblance. The Seminar of Jacques La- can, Book XVIII. Translated by Cormac Gallagher from unedited French manuscripts, http://www.lacaninireland.com, (accessed 07.04.17) Seminar in France (1971), French edi- tion, Editions du Seuil, 2007. [21] J. Lacan, O Semin´ario, Livro 8: A Transferencia, Jorge Zahar, Rio de Janeiro, 1992, First French edition in 1991. [22] E. Tagliazucchi, P. Balenzuela, D. Fraiman, D. R. Chialvo, Criticality in large-scale brain fMRI dynamics unveiled by a novel point process analysis, Frontiers in Physiology -- Fractal Physiology 3, (2012) Article 15. [23] E.R. Kandel, J.H. Schwartz, T.M. Jessel (Eds.), Principles of Neural Science, MacGraw Hill, New York, 2000. [24] H. Hartline, F. Ratcliff, Inhibitory Interactions of Receptor Units in the Eye of Limulus, Journal of General Physiology 40 (1957) 357 -- 376. [25] W.F. Ganong, Review of Medical Physiology, MacGraw Hill, USA, 2003. [26] L.A.V. de Carvalho, D.Q. Mendes, R.S. Wedemann, Creativity and Delusions: The Dopamin- ergic Modulation of Cortical Maps, Lecture Notes in Computer Science 2657 (2003) 511 -- 520. [27] J.A. Hertz, A. Krogh, R.G. Palmer, Introduction to the Theory of Neural Computation, Perseus Books, Cambridge, Massachusetts, USA, 1991. [28] G.M. Edelman, Wider than the Sky, a Revolutionary View of Consciousness, Penguin Books, London, 2005. [29] E.R. Kandel, I. Kupfermann, S. Iversen, Learning and Memory, in: E.R. Kandel, J.H. Schwartz, T.M. Jessel (Eds.), Principles of Neural Science, fourth ed., MacGraw Hill, New York, 2000, pp. 1227-1246. [30] W.C. Barbosa, Massively Parallel Models of Computation, Ellis Horwood Limited, England, 23 1993. [31] R.S. Wedemann, L.A.V. de Carvalho, R. Donangelo, Access to Symbolization and Associativ- ity Mechanisms in a Model of Conscious and Unconscious Processes, in: A. V. Samsonovich and K. R. J´ohannsd´ottir (Eds.), Biologically Inspired Cognitive Architectures, Frontiers in Artificial Intelligence and Applications, IOS Press, USA, 2011, pp. 444 -- 449. [32] S.A. Huettel, A.W. Song, G. McCarthy, Functional Magnetic Resonance Imaging, second ed., Sinauer, Massachusetts, USA (2009). [33] J.M. Beggs, D. Plenz, Neuronal avalanches in neocortical circuits, Journal of Neuroscience 23 (2003) 11167 -- 11177. [34] H.J. Jensen, Self-Organized Criticality, Emergent Complex Behavior in Physical and Biological Systems, Cambridge University Press, Cambridge, UK, 1998. [35] V. Capano, H.J. Herrmann, L. de Arcangelis, Optimal percentage of inhibitory synapses in multi-task learning, Nature Sci. Reps. 5 (2015) Article 9895. [36] R.S. Wedemann, A.R. Plastino, Asymmetries in Synaptic Connections and the Nonlinear Fokker-Planck Formalism. Lecture Notes in Computer Science 9886 (2016) 19 -- 27. 24
1608.01179
1
1608
2016-08-03T13:14:05
Normalized neural representations of natural odors
[ "q-bio.NC", "physics.bio-ph" ]
The olfactory system removes correlations in natural odors using a network of inhibitory neurons in the olfactory bulb. It has been proposed that this network integrates the response from all olfactory receptors and inhibits them equally. However, how such global inhibition influences the neural representations of odors is unclear. Here, we study a simple statistical model of this situation, which leads to concentration-invariant, sparse representations of the odor composition. We show that the inhibition strength can be tuned to obtain sparse representations that are still useful to discriminate odors that vary in relative concentration, size, and composition. The model reveals two generic consequences of global inhibition: (i) odors with many molecular species are more difficult to discriminate and (ii) receptor arrays with heterogeneous sensitivities perform badly. Our work can thus help to understand how global inhibition shapes normalized odor representations for further processing in the brain.
q-bio.NC
q-bio
Normalized neural representations of natural odors David Zwicker1, 2, ∗ 1School of Engineering and Applied Sciences, Harvard University, Cambridge, MA 02138, USA 2Kavli Institute for Bionano Science and Technology, Harvard University, Cambridge, MA 02138, USA (Dated: July 3, 2018) The olfactory system removes correlations in natural odors using a network of inhibitory neurons in the olfactory bulb. It has been proposed that this network integrates the re- sponse from all olfactory receptors and inhibits them equally. However, how such global inhibition influences the neural representations of odors is unclear. Here, we study a simple statistical model of this situation, which leads to concentration-invariant, sparse representations of the odor composition. We show that the inhibition strength can be tuned to obtain sparse representations that are still useful to discriminate odors that vary in relative concentration, size, and composition. The model reveals two generic consequences of global inhibition: (i) odors with many molecular species are more dif- ficult to discriminate and (ii) receptor arrays with heterogeneous sensitivities perform badly. Our work can thus help to understand how global inhibition shapes normalized odor representations for further processing in the brain. I. INTRODUCTION Sensory systems encode information efficiently by re- moving redundancies present in natural stimuli (Barlow, 2001, 1961). In natural images, for instance, neighbor- ing regions are likely of similar brightness and the image can thus be characterized by the regions of brightness changes (Ruderman and Bialek, 1994). This structure is exploited by ganglion cells in the retina that respond to brightness gradients by receiving excitatory input from photo receptors in one location and inhibitory input from the surrounding (Demb and Singer, 2015). This typi- cal center-surround inhibition results in neural patterns that represent natural images efficiently (Carandini and Heeger, 2012). Similarly, such local inhibition helps sepa- rating sound frequencies in the ear and locations touched on the skin (Isaacson and Scanziani, 2011). Vision, hear- ing, and touch have in common that their stimulus spaces have a metric for which typical correlations in natural stimuli are local. Consequently, local inhibition can be used to remove these correlations and reduce the high- dimensional input to a lower-dimensional representation. The olfactory stimulus space is also high-dimensional, since odors are comprised of many molecules at different concentrations. Moreover, the concentrations are also often correlated, e. g., because the molecules originate from the same source. However, these correlations are not represented by neighboring neurons in the olfactory system, since there is no obvious similarity metric for molecules that could be used to achieve such an arrange- ment (Nikolova and Jaworska, 2003). Because the olfac- tory space lacks such a metric, local inhibition cannot be used to remove correlations to form an efficient represen- tation (Murthy, 2011; Soucy et al., 2009). Consequently, ∗ http://www.david-zwicker.de the experimentally discovered inhibition in the olfactory system (Yokoi, Mori, and Nakanishi, 1995) likely affects neurons irrespective of their location. Such global inhi- bition could for instance normalize the activities by their sum, which has been observed experimentally (Olsen, Bhandawat, and Wilson, 2010; Roland et al., 2016). This normalization cannot reduce the correlation structure of odors, but it could help separating the odor composition (what is present?) from the odor intensity (how much is there?) (Cleland, 2010; Laurent, 1999). This separation is useful, since the composition identifies an odor source, while the intensity information is necessary for finding or avoiding it. However, how global inhibition shapes such a bipartite representation of natural odors is little understood. In this paper, we study a simple model of the ol- factory system that resembles its first processing lay- ers, which transform the odor representation succes- sively (Silva Teixeira, Cerqueira, and Silva Ferreira, 2016; Wilson, 2013), see Fig. 1. Our model connects previous results from simulations of the neural circuits (Cleland and Sethupathy, 2006; Getz and Lutz, 1999; Li, 1990, 1994; Linster and Hasselmo, 1997; Zhang, Li, and Wu, 2013) to system-level descriptions of the olfactory sys- tem (Hopfield, 1999; Koulakov, Gelperin, and Rinberg, 2007; Zwicker, Murugan, and Brenner, 2016). The main feature of the model is global inhibition, which leads to normalization. This separates the odor composition from its intensity and encodes it in a sparse representation. The inhibition strength controls the trade-off between the sparsity and the transmitted information, which in- fluences how well this code can be used to discriminate odors in typical olfactory tasks. The model reveals two generic consequences of global inhibition: (i) odors com- prised of many different molecules exhibit sparser rep- resentations and should thus be more difficult to distin- guish and (ii) overly sensitive receptors could dominate arXiv:1608.01179v1 [q-bio.NC] 3 Aug 2016 2 FIG. 1 Schematic picture of our model describing the signal processing in the olfactory bulb: An odor comprised of many ligands excites the olfactory receptors and the signals from all receptors of the same type are accumulated in respective glomeruli. Associated projections neurons receive excitatory input from a single glomerulus and are subject to global inhibition, mediated by a network of local neurons. The activity of the projection neurons form a sparse, concentration-invariant odor representation. the sparse responses and arrays with heterogeneous re- ceptors should thus perform poorly. II. SIMPLE MODEL OF THE OLFACTORY SYSTEM Odors are blends of odorant molecules that are lig- ands of the olfactory receptors. We describe an odor by a vector c = (c1, c2, . . . , cNL ) that specifies the concen- trations ci of all NL detectable ligands (ci ≥ 0). Gen- erally, only a small subset of the NL ∼ 105 ligands are present in natural odors, so most of the ci will typically be zero. The ligands in an odor are detected by olfactory receptor neurons, which reside in the nose in mammals and in the antenna in insects (Kaupp, 2010). Each of these neurons expresses receptors of one of NR genet- ically defined types, where NR ≈ 50 for flies (Wilson, 2013), NR ≈ 300 for humans (Verbeurgt et al., 2014), and NR ≈ 1000 for mice (Niimura, 2012). The excitation of all receptor neurons of the same type is accumulated in associated glomeruli (Su, Menuz, and Carlson, 2009), whose excitation pattern forms the first odor representa- tion, see Fig. 1. Here, the large number of ligands and their possible mixtures are represented by a combina- torial code, where each ligand typically excites multiple receptor types (Malnic et al., 1999). It has been shown experimentally that the excitation en of the glomerulus associated with receptor type n can be approximated by a linear function of the ligand concentrations c (Gupta, Albeanu, and Bhalla, 2015; Silbering and Galizia, 2007; Tabor et al., 2004), en = NLXi=1 Snici , (1) where Sni denotes the sensitivity of glomerulus n to lig- and i. We here consider a statistical description of com- binatorial coding by studying random sensitivity matri- ces with entries drawn independently from a log-normal distribution. This distribution is parameterized by the mean sensitivity ¯S and the standard deviation λ of the underlying normal distribution. This choice is moti- vated by experimental measurements, which also suggest that λ ≈ 1 for flies and humans (Zwicker, Murugan, and Brenner, 2016). We showed previously that such random matrices typically decorrelate stimuli and thus lead to near-optimal odor representations on the level of glomeruli (Zwicker, Murugan, and Brenner, 2016). In contrast to our previous model, we here consider the odor representation encoded by projection neurons (mitral and tufted cells in mammals), which constitute the next layer after the glomeruli, see Fig. 1. Projec- tion neurons typically receive excitatory input from a single glomerulus (Jefferis et al., 2001) and inhibitory input from many local neurons (granule cells in mam- mals), which are connected to other projection neurons and glomeruli (Cleland, 2010; Su, Menuz, and Carlson, 2009). The activity an of the projection neurons associ- ated with receptor type n is a sigmoidal function of lig- and concentrations (Bhandawat et al., 2007; Tan et al., 2010). Additionally, all signals are subject to noise, both from stochastic ligand-receptor interactions and from in- ternal processing (Lowe and Gold, 1995), which limits the number of distinguishable output activities. We capture both effects by considering the simple case where only two activities an can be distinguished. Here, the projec- tion neurons are active when their excitatory input, the respective excitation en, exceeds a threshold γ, an =(0 1 en ≤ γ en > γ . (2) Generally, γ could depend on the type n, but we here consider a simple mean-field model, where all types ex- hibit the same threshold. Nevertheless, this threshold could still depend on global variables. Experimental data (Asahina et al., 2009; Aungst et al., 2003; Banerjee et al., 2015; Berck et al., 2016; Hong and Wilson, 2015; Olsen, Bhandawat, and Wilson, 2010; Roland et al., 2016; Silbering and Galizia, 2007) and modeling of the local neurons (Cleland, 2010; Cleland and Sethupathy, 2006) suggest that the total excitation of all glomeruli inhibits all projection neurons. To capture this we postulate that Nose (antenna) Receptors Olfactory bulb (antennal lobe) Glomeruli Projection neurons Olfactory cortex (mushroom body) composition intensity Local neurons YY Y YYY Y Odor c Excitation e Activity a 3 is because both the excitations en and the threshold γ are linear in c and Sni, see Eqs. 1 and 3, and the ac- tivities an only depend on the ratio en/γ, see Eq. 2. In fact, these equations can be interpreted as normaliza- tion of the excitations by the total excitation followed by thresholding with the constant threshold α/NR. Since the representation a does not depend on ctot, it only en- codes relative ligand concentrations, i. e., the odor com- position. This property is called concentration invariance and corresponds to the everyday experiences that odors smell the same over many orders of magnitude in concen- tration (Cleland et al., 2011; Uchida and Mainen, 2007; Zhang, Li, and Wu, 2013). Indeed, experiments suggest that the activity of projection neurons is concentration- invariant (Cleland et al., 2007; Sachse and Galizia, 2002; Sirotin, Shusterman, and Rinberg, 2015) and exhibits more uniform distances between odors (Bhandawat et al., 2007; Cleland et al., 2007), indicating that they encode the odor composition efficiently. To understand how odor compositions are encoded in our model, we start with numerical simulations of Eqs. 1– 3 as described in the SI. Fig. 2A shows the excitations en corresponding to an arbitrary odor. Here, the excitation threshold is 1.4 times the mean excitation, and only three channels are active (orange bars). The corresponding his- togram in Fig. 2B shows that the number of active chan- nels is typically small for this inhibition strength when odors are presented with statistics Penv(c). Moreover, the magnitude of the Pearson correlation coefficient be- tween two channels is typically only 1 %, see SI. This weak correlation is expected for the uncorrelated odors and random sensitivity matrices that we consider here and explains why the histogram in Fig. 2B is close to a binomial distribution. The odor representations are thus mainly characterized by the mean channel activity hani. The mean channel activity hani depends on the in- hibition strength α, the sensitivities Sni, and the odor statistics Penv(c). To discuss these dependences, we next introduce an approximation based on a statistical de- scription of the associated excitation en. Here, we define the normalized concentrations ci = ci/ctot and normal- ized excitations en = en/(ctot ¯S), since an is independent of ctot and ¯S. The statistics of ci can be estimated in the typical case where odors are comprised of many lig- ands, see SI. In the particular case where the ligands are identically distributed the mean is hcii = NL−1 and the 2). Gen- variance reads var(ci) ≈ (1 − p + σ2µ−2)/(pNL erally, ci varies more if the underlying ci has higher coef- ficient of variation σi/µi or if the mean odor size s = pNL is smaller. The normalized excitation en is de- fined such that its mean is 1 and the associated variance can be written as a product of the external contribution Vext =Pihc2 ii due to odors and the internal contribution Vint = var(Sni)hSnii−2 due to sensitivities, see SI. In the the threshold γ is a function of the total excitation, where we for simplicity consider a linear dependence, (3) en . NRXn=1 R α N γ = Here, α is a parameter that controls the inhibition strength. Taken together, our model of the olfactory system com- prises NR communication channels, each consisting of receptors, a glomerulus, and projection neurons, which interact via global inhibition, see Fig. 1. The Eqs. 1–3 describe how this system maps an odor c to an activ- ity pattern a = (a1, a2, . . . , aNR). The amount of infor- mation that can be learned about c by observing a is quantified by the mutual information I, which reads I = −Xa P (a) log2 P (a) . (4) Here, the probability P (a) of observing output a is given by P (a) = R P (ac)Penv(c) dc. The conditional proba- bility P (ac) of observing a given c describes the process- ing in the olfactory system and follows from the Eqs. 1–3. In contrast, Penv(c) denotes the probability of encoun- tering an odor c, which depends on the environment. Consequently, the information I is not only a function of the sensitivity matrix Sni and the inhibition strength α, but also of the environment in which the receptors are used (Zwicker, Murugan, and Brenner, 2016). Natural odor statistics are hard to measure (Wright and Thomson, 2005) and we thus cannot infer the distri- bution Penv(c) from experimental data. Instead, we con- sider a broad class of distributions parameterized by a few parameters. For simplicity, we only consider uncorrelated odors, where the concentrations ci of ligands are indepen- dent. We denote by pi the probability that ligand i is part of an odor. If this is the case, the associated ci is drawn from a log-normal distribution with mean µi and stan- dard deviation σi. This choice allows us to independently adjust the mean odor size s = Pi pi, the mean of the total concentration ctot = Pi ci, and the concentration variations σi . Averaged over all odors, ci then has mean µi i + piσ2 hcii = piµi and variance var(ci) = (pi − p2 i . Note that typical odors can have hundreds of different ligands (Wright and Thomson, 2005), but this is still well below NL ∼ 105 and we thus have 1 (cid:28) s (cid:28) NL. i )µ2 III. RESULTS A. Global inhibition leads to concentration-invariant, sparse representations Our model has the interesting property that the odor representation a does not change when the odor c or the sensitivities Sni are scaled by a positive factor. This 4 FIG. 2 Global inhibition with thresholding leads to sparse odor representations a. (A) Excitations en for an arbitrary odor. Active channels (orange) have an excitation above the threshold (red line, inhibition strength α = 1.4). The right axis indicates the normalized excitation en = enNR/Pm em. (B) Histogram of the number of active channels compared to a binomial distribution (black line) with the same mean for α = 1.4. (C) Mean channel activity hani as a function of α. The approximation given by Eq. 6 (solid line) is compared to numerical simulations (symbols, standard error of the mean smaller than symbol size). The gray dotted line indicates a single expected active channel in humans, hani = 1 300 . (A–C) Additional model parameters are NR = 32, NL = 256, pi = 0.1, µi = σi = 1, and λ = 1. receptor sensitivities, which determine Vext and Vint, re- spectively. In particular, the width λ of the sensitiv- ity distribution could also be under evolutionary control. However, experimental data suggests that both flies and humans exhibit λ ≈ 1 (Zwicker, Murugan, and Brenner, 2016). Additionally, we show in the SI that much smaller or larger values lead to extremely sparse representations, such that we will only consider λ = 1 in the following. In this case, the inhibition strength α controls the spar- sity of the odor representation in our simple model of the olfactory system. B. Sparse coding transmits useful information One problem with sparse representations is that they cannot encode as many odors as dense representations. There is thus a maximal sparsity at which typical olfac- tory tasks can still be performed. In general, the perfor- mance of the olfactory system can be quantified by the transmitted information I, which is defined in Eq. 4. If we for simplicity neglect the small correlations between channels, I can be approximated as (Zwicker, Murugan, and Brenner, 2016) I ≈ − NRXn=1(cid:2)hani log2hani + (1 − hani) log2(1 − hani)(cid:3) . (7) A maximum of NR bits is transmitted when half the chan- nels are active on average, hani = 1 2 . In our model, this is the case for weak inhibition, α < 1, see Fig. 2C. In the opposite case of significant inhibition, α > 1, few chan- nels are typically active and the transmitted information is smaller. In the limit hani (cid:28) 1, the information is ap- proximately given by I ∼ 1 ln 2 NRhani· (1− lnhani), which implies that even if only 10 % of the channels are active on average, the information I is still almost half of the max- imal value of NR bits. However, large information I does not automatically indicate a good receptor array, since only accessible information that can be used to solve a − 1 , (5) (cid:18)1 + σ2 µ2(cid:19) Vint = eλ2 1 s var(en) = VextVint Vext ≈ simple case of identically distributed ligands, we have for 1 (cid:28) s (cid:28) NL, see SI. The normalized excitations thus vary more if odors contain fewer ligands, concentra- tions fluctuate stronger, or sensitivities are distributed more broadly. Finally, the mean channel activity hani is given by the probability that the excitation en exceeds the threshold γ, see Eq. 2. This is equal to the probability that the normalized excitation en exceeds the normalized threshold γ = γ/( ¯Sctot). Replacing γ by its expectation value hγi = α and using log-normally distributed en, we obtain ln(cid:0)1 + VextVint(cid:1)(6) 1 2 erfc(cid:18) ζ + ln α 2 (cid:19) with ζ = 2ζ 1 1 2 hani ≈ for log-normally distributed en, see SI. Fig. 2C shows that this is a good approximation of the numerical re- sults, which have been obtained from ensemble averages of Eq. 2. The mean activity hani can also be interpreted as the mean fraction of channels that are activated by an odor, such that small hani corresponds to sparse odor repre- sentations. Fig. 2C shows that in our model this is the case for large inhibition strength α, where hani ∼ e−ν with ν ≈ (ln α)2/(4ζ), see SI. Since sparse representa- tions are thought to be efficient for further processing in the brain (Laurent, 1999; Olshausen and Field, 2004) the inhibition strength α could be tuned, e. g., on evo- lutionary time scales, to achieve an activity hani that is optimal for processing the odor representation down- stream. If the optimal value of hani is the same across animals, our theory predicts that inhibition is stronger in systems with more receptor types. However, this sim- ple argument is not sufficient, since hani also depends on the variations in the natural odor statistics and the Numerics Approximation, Eq. 6 0 1 2 Inhibition strength α 1 0.1 0.01 Activity(cid:31)an(cid:30) C 0 5 Active channel count(cid:31)n an 0.3 0.2 0.1 0 B Frequency 2 1 0 5 10 15 20 Channel n 25 30 30 25 20 15 10 05 A Excitationen 5 considered. This qualitatively agrees with experiments where humans are not able to identify all ligands in mix- tures of more than three ligands (Goyert et al., 2007; Jinks and Laing, 2001) and they fail to detect the pres- ence or absence of ligands in mixtures of more then 15 ligands (Jinks and Laing, 1999). Even if humans cannot identify ligands in large odors, they might still be able to distinguish two such odors. To study this, we next compare the representations of two odors that each contain s ligands, sharing sB of them, for the simple case where all ligands have the same con- centration. Fig. 3C shows that the distance hdi between the two odors decreases with larger sB, i. e., more sim- ilar odors are more difficult to discriminate. However, sB only has a strong effect if more than about 80 % of the ligands are shared between odors. Conversely, the inhibition strength α and the odor size s significantly influence hdi for all values of sB. This agrees with the results shown in Fig. 3B, where hdi exhibits a similar de- pendence on α and s. While it is expected that the per- formance decreases with large inhibition strength α since fewer channels are active, the strong dependence on the size s is surprising. Why are odors with many ligands more difficult to dis- criminate in our model? Since correlations between chan- nels seem to be negligible, the most likely explanation is that larger odors activate fewer channels. To test this hypothesis, we determine the activity hani in the simple case where all ligands in an odor have the same concen- tration. Because of the normalization, the value of this concentration does not matter and hani only depends on In the the inhibition strength α and the odor size s. limit of large odors (s (cid:29) 1), the approximation given in Eq. 6 yields hani ∼ e−βs with β ∼ (ln α)2, see SI. In this case, the activity hani thus decreases exponentially with s and this decrease is stronger for larger α. Con- sequently, larger odors activate fewer channels and it is thus less likely that a small change in such odors alters the activation pattern a. Larger odors activate fewer channels because the re- spective excitations en have a smaller variability. For an odor with s ligands of equal concentration, en is pro- portional to the sum of s sensitivities Sni, see Eq. 1. Consequently, en can be considered as a random vari- able whose mean heni and variance var(en) scale with s. The activity hani is given by the fraction of exci- tations that exceed the threshold γ, which also scales with s. This fraction typically scales with the coefficient heni−1, which is proportional to s− 1 of variation var(en) 2 and is thus smaller for larger odors. Larger odors thus activate fewer channels because there are fewer excita- tions that are much larger than the mean, see Fig. 4A. 12 C. Larger odors have sparser representations given task matters (Tikhonov, Little, and Gregor, 2015; Tkacik and Bialek, 2016). To test whether sparse representations are sufficient to solve typical olfactory tasks, we next study how well odors can be discriminated in our model. As a proxy for the discriminability, we calculate the Hamming dis- tance d between the odor representations, which is given by the number of channels with different activity. In the simple case of uncorrelated odors, which do not share any ligands, the expected distance hdi is approximately given by total number of active channels in both repre- sentations. Consequently, uncorrelated odors can be dis- tinguished even if their representations are very sparse. However, realistic tasks typically require distinguishing similar odors. We thus next study the discriminability of odors that vary in the relative concentrations of their ligands, their size, and their composition. ct cb We start by determining the maximal dilution cb ct at which a target odor at concentration ct can still be de- tected in a background of concentration cb. We calcu- late the expected difference hdi between the associated representations from the probability that a given chan- nel changes its activity when the target is added, see SI. Since this probability is the same for all channels, hdi is proportional to the number NR of channels. For the sim- ple case where both the target and the background are a single ligand, Fig. 3A shows that hdi decreases for smaller target concentrations and is qualitatively the same for all inhibition strengths α. For large dilutions cb , hdi is in- ct versely proportional to the dilution, hdi ∝ NR . Since the addition of the target can only be detected reliably if hdi > 2, which corresponds to a situation where one chan- nel becomes inactive and another one active, our model predicts that doubling the number NR of channels also doubles the concentration sensitivity. Fig. 3A thus im- plies that mice (NR ≈ 1000) should be able to detect the addition of a target even if it is almost a hundred times more dilute than the background, which is close to the threshold that has been found experimentally (Mouret et al., 2009). Conversely, flies (NR ≈ 50) should fail for very small dilution factors. We next study odors comprised of many ligands, since typical odors are blends (Wright and Thomson, 2005). For simplicity, we consider the detection of a single tar- get ligand in a background mixture of varying size s when the target ligand and the ligands in the background have equal concentration, such that the target dilution is s. Fig. 3B shows that the qualitative dependence of hdi on the dilution is similar to the single ligand case in panel A, but the maximal dilution for detecting the target is dif- ferent. For instance, the model predicts that mice cannot identify the addition of the target ligand to a background consisting of more than ten ligands, while the maximal di- lution was almost one hundred in the case of single back- ground ligands. Consequently, the discrimination perfor- mance seems to drop significantly when larger odors are 6 FIG. 3 Sparse coding is sufficient to distinguish odors with different relative ligand concentrations, size, and composition. (A) Mean distance hdi between the representations of a background ligand at concentration cb and an odor with an additional (B) Distance hdi target ligand at concentration ct as a function of the dilution cb/ct for various inhibition strengths α. resulting from adding a ligand to an odor comprised of s ligands as a function of s for various α. (C) Distance hdi between the representations of two odors with s ligands, sharing sB of them, as a function of the similarity sB/s for small (s = 8, solid lines) and large odors (s = 16, dashed lines). The colors indicate the same α as in the other panels. (A–C) The gray dotted lines indicate the threshold hdi = 2 for NR = 50, 300, 1000 (corresponding to flies, humans, and mice; top to bottom). The width of the sensitivity distribution is λ = 1. FIG. 4 Larger odors activate fewer channels. (A) Comparison of the excitations en of small (odor size s = 8, upper panels) and large odors (s = 64, lower panels) at α = 1.3. en for a single odor (left panels) and histograms for all odors (right panels) are shown. Larger odors exhibit fewer active channels (dark blue), for which the excitations are above threshold (red line). (B) Numerically determined hani as a function of s for various inhibition strengths α at small (σ/µ = 1, solid lines) and large concentration variability (σ/µ = 10, dashed lines) at NL = 104. (C) Numerically determined hani as a function of σ/µ for various α. (A–C) Additional model parameters are NR = 32, NL = 256, pi = 0.1, µi = σi = 1, and λ = 1. The gray dotted line in B and C indicates a single expected active channel in humans, hani = 1 300 . This is a direct consequence of the assumption that the excitation threshold γ scales with the mean excitation and this result does not depend on other details of the model. Conversely, the dependence of hani on the inhi- bition strength α is model specific, since it follows from the shape of the tail of the excitation distribution. In particular, the influence of the odor size on hani is in- significant for weak inhibition, α ≈ 1, because approxi- mately half the channels are activated irrespective of the variance var(en). This qualitative explanation illustrates that depend- ing on the variability of the excitations different odors can have representations with very different sparsities. Indeed, we find that the sparsity changes over several or- ders of magnitude as a function of the odor size s in our model, see Fig. 4B. Moreover, the concentration variabil- ity σ µ of the individual ligands also has a strong effect on the sparsity, see Fig. 4C. This is because larger σ µ implies larger variations in the excitations, such that more chan- nels exceed the threshold and become active. In fact, this dependence of hani on s and σ µ is also qualitatively captured by the analytical approximation given in Eq. 6, which explicitly depends on the odor variability Vext de- fined in Eq. 5. Taken together, our model shows that the sparsity of the odor representations strongly depend on the odor statistics Penv(c). D. Effective arrays have similar receptor sensitivities So far, we considered homogeneous receptor arrays, where all receptor types have the same average sensitiv- ity. However, realistic receptors vary in their biochem- ical details and it might thus be difficult to have such homogeneous arrays. We thus next consider the effect of sensitivity variations between different receptors. This is important, since a channel with overly sensitive receptors will contribute significantly to the common threshold γ, suppress the activity of other channels, and could thus limit the coding capacity of the system, see Fig. 5A. To study this, we consider sensitivity matrices Sni = ξnSiid ni , where ξn denotes the mean sensitivity of receptor type n and Siid is the sensitivity matrix that we discussed so ni far, i. e., it is a random matrix where all entries are in- dependently drawn from a log-normal distribution de- scribed by the mean ¯S and width λ. Here, ξn captures differences between receptor types, e. g., because of bio- chemical differences or due to variations in copy number, 1 100 Concentration ratio cb/ct 10 1 10 Background mixture size s 0 0.2 0.8 Composition similarity sB/s 0.4 0.6 A Relative ligand concentration B Odor size C Odor composition s = 8 s = 16 0.1 0.01 10−3 Dist.(cid:31)d(cid:30)/NR 0.1 0.01 α = 1.5 α = 2 α = 3 10−3 Dist.(cid:31)d(cid:30)/NR α = 1.5 α = 2 α = 3 Fly Human Mouse 0.1 0.01 10−3 Dist.(cid:31)d(cid:30)/NR α = 1.5 α = 2 α = 3 0.1 Concentration variability σ/µ 10 1 1 0.1 0.01 10−3 10−4 10−5 Activity(cid:31)an(cid:30) C α = 1.5 α = 2 α = 3 1 100 10 Mixture size s 1000 1 0.1 0.01 10−3 10−4 10−5 Activity(cid:31)an(cid:30) B 20 10 Channel n 30 0.1 0 Frequency 15 10 05 100 50 0 A en en see SI. For this model, the mean excitation threshold is hγi = α ¯ShctotiξtotNR−1 where ξtot = Pn ξn. The ex- pected channel activity is approximately given by Nrξn(cid:19) , hani ≈ 1 − F(cid:18) αξtot (8) where F (en) is the cumulative distribution function of the normalized excitations en for ξn = 1, whose mean is heni = 1 and whose variance is given by Eq. 5. Note that hani does not change if all ξn are multiplied by the same factor. In particular, the expression above reduces to hani ≈ 1 − F (α) and thus Eq. 6 if all ξn are equal. We first discuss the influence of the receptor sensitivi- ties ξn by only varying one type, i. e., we change ξ1 while setting ξn = 1 for n ≥ 2. Fig. 5B shows that for fixed channel activity hani the transmitted information I is maximal for a homogeneous receptor array (ξ1 = 1). I is reduced for smaller ξ1 and for ξ1 = 0 it reaches the value I0 of an array where the first receptor was removed. Conversely, I can drop well below I0 when ξ1 is increased above 1. In this case, the large excitation of the affected channel not only leads to its likely activation, but it also raises the threshold γ and thereby inhibits other chan- nels, see Fig. 5A. In the extreme case of very large ξ1, this channel will always be active while all other chan- nels are silenced, which implies I = 0. There is thus a critical value of ξ1 beyond which removing the receptor from the array is advantageous for the overall perfor- mance. Fig. 5B shows that increasing the sensitivity of a receptor by only 40 % can make it useless in the context of the whole array if representations are sparse. So far, we only varied the sensitivity of a single re- ceptor. To test how variations in the sensitivities of all receptors affect the information I, we next consider log- normally distributed ξn. Here, vanishing variance of ξn corresponds to a homogeneous receptor array. Fig. 5C shows that small variations in ξn can strongly reduce the transmitted information I. Since I limits the discrimi- native capability of the receptor array, this suggests that receptor arrays with heterogeneous sensitivities perform worse. The simple model that we discuss here shows that the excitation statistics of the different channels deter- mine the properties of the resulting odor representation. In particular, receptors that have lower excitations on average might be suppressed often and thus contribute less to the odor information. Since the excitation statis- tics are influenced both by the sensitivities Sni and the odor statistics Penv(c), this suggests that the sensitivities should be adjusted to the odor statistics. In an optimal receptor array, the sensitivities are chosen such that all channels have the same probability to become active. 7 IV. DISCUSSION We studied a simple model of odor representations, which is based on normalization and a non-linear gain function. This model separates the odor composition, encoded in the activity a of the projection neurons, from the odor intensity, which could be encoded by the to- tal excitation etot or the threshold level γ (Mainland et al., 2014). For significant inhibition the representa- tion a is sparse and the set of active projection neurons provides a natural odor 'tag' that could be used for iden- tification and memorization in the downstream process- ing (Stevens, 2015). Sparse representations reduce the coding capacity and transmit less information than dense ones. However, even if the mean activity is hani = 0.01 and thus 50 times smaller than in maximally informative arrays with hani = 0.5, the transmitted information I is only reduced by a factor of 12, see Eq. 7. For humans with NR = 300, this yields I ≈ 25 bits, allowing to encode 2I ≈ 107 dif- ferent odor compositions. Note that the total informa- tion Itot also includes information Iint about the odor intensity, Itot = I + Iint. Here, Iint ≈ 10 bits would be sufficient to encode the total concentration over a range of 10 orders of magnitude with a resolution of 5 %, typ- ical for humans (Cain, 1977). In this case, our model compresses the 300 bits of a maximally informative rep- resentation on the level of glomeruli (Zwicker, Murugan, and Brenner, 2016) to only Itot ≈ 35 bits on the level of projection neurons. The model discussed here is similar to our previous model, where we discussed representations on the level of the glomeruli (Zwicker, Murugan, and Brenner, 2016). Both models use a maximum entropy principle to deter- mine properties of optimal receptor arrays. To achieve this, the receptor sensitivities must be tailored to the odor statistics in both models. The main difference of the models is the global inhibition discussed here, which sep- arates the odor composition from its intensity and thus removes the correlation between the glomeruli excitation and the odor intensity (Haddad et al., 2010). Conse- quently, odors can then be discriminated at all concen- trations, while this was only possible in a narrow concen- tration range in the glomeruli model (Zwicker, Murugan, and Brenner, 2016). The additional normalization is thus useful to separate odors, even if the projection neurons encode less information than the respective glomeruli. To estimate this information, we consider binary out- puts in both models, which corresponds to very noisy channels. However, the glomeruli model discusses arrays of noisy receptor, while we here consider perfect recep- tors whose signal is first normalized and then subjected to noise. This additional processing reduces correlations and leads to sparse representations, which might simplify downstream computations. Consequently, this model is suitable for describing natural olfaction, where the capac- 8 FIG. 5 Receptors with heterogeneous sensitivities make poor arrays. (A) Comparison of the excitations en for homogeneous (ξ1 = 1, upper panels) and heterogeneous receptors (ξ1 = 2, lower panels). en for the same arbitrary odor (left panels) and histograms for all odors (right panels) are shown for the first receptor (n = 1, orange) and all other receptors (n ≥ 2, blue). Dark bars indicate excitations that are above the threshold (red line, inhibition strength α = 1.3). (B) Information I given by Eq. 7 as a function of the sensitivity ξ1 of the first receptor. The channel activity hani calculated from Eq. 8 is set to the given value by adjusting α. I is shown relative to the information I0 of a system without the first receptor (dotted line). (C) Information I (line, mean; shaded area indicates standard deviation) of log-normally distributed ξn as a function of the variation var(ξn)hξni−2 for various hani. (A–C) Remaining parameters are NR = 32, NL = 256, pi = 0.1, µi = σi = 1, and λ = 1. ity for the downstream computations is limited, while the glomeruli model is relevant for artificial olfaction (Stitzel, Aernecke, and Walt, 2011), since computers have enough power to handle high-dimensional signals. Sparse responses of projection neurons have been ob- served in experiments (Davison and Katz, 2007; Rinberg, Koulakov, and Gelperin, 2006). For instance, in mice 15 % of the projection neurons respond to a given single ligand (Roland et al., 2016), suggesting significant inhi- bition. However, in locust about two third of the pro- jection neurons respond to any given odor (Perez-Orive et al., 2002), which implies weak inhibition. It is thus conceivable that some animals exhibit sparse representa- tions while others have maximally informative ones, al- though additional experiments are needed to character- ize the representations better. A direct experiment could test whether the odor percept changes when the weakly responding glomeruli are disabled artificially. Addition- ally, it will be important to study the representations of mono-molecular odors and mixtures at various concentra- tion to better resemble the natural odor statistics. For instance, our simple theory predicts that fewer than 15 % of the projection neurons in mice respond when complex mixtures are presented. Indeed, experiments find that only 3 to 10 % of the projection neurons in mice fire for complex urine odors (Lin et al., 2005). Conversely, the statistics of the activity of projection neurons in flies seem to be independent of the stimulus (Stevens, 2016). Our theory can also be tested by measuring how well odors can be discriminated. For instance, odors are much more difficult to distinguish if they contain more ligands in our model, which has also been observed experimen- tally (Weiss et al., 2012). Conversely, other experiments indicate that the odor size only weakly influences the odor discriminability (Bushdid et al., 2014). Taken to- gether, there is some experimental evidence that the odor representations and thus the discriminability change with odor size, although there is also evidence to the contrary, which could hint at mechanisms beyond global inhibition that influence the odor representations. The coding sparsity given by the mean channel ac- tivity hani can be adjusted by changing the inhibition strength α or the width λ of the receptor sensitivity dis- tribution in our model. Additionally, hani is a function of the natural odor statistics, i. e., the typical number of ligands in odors and their concentration distribution. Consequently, α or λ must be adjusted to keep hani con- stant if the odor statistics change, e. g., because of sea- sonal changes or migration to a different environment. This adjustment could happen on multiple timescales, reaching from evolutionary adaptations of the receptors to near-instantaneous adjustments of the involved neu- rons, and it is likely that the global inhibition is reg- ulated on all levels (Wilson, 2013). In this paper, we investigated the simple case of constant α and λ, which corresponds to slow regulation, but it is conceivable that α could be regulated on short time scales. For instance, the threshold could be lowered for larger odors to im- prove their discriminability. Our model suggests that such additional mechanisms are necessary to efficiently discriminate odors of all sizes. Our model also reveals that it is important to control the properties of the individual communication channels to have useful receptor arrays. For instance, increasing the sensitivity of a given receptor by 40 % can be worse then removing it completely, see Fig. 5A. Generally, a receptor array is only effective if the different channels have similar excitations on average. This suggests that the sensitivities are tightly controlled and maybe even adjusted to the odor statistics of the environment. On evolutionary time scales, the sensitivities could be regu- lated by point mutations of the receptors that change how ligands bind (Yu et al., 2015). On shorter time scales, the sensitivities could be regulated by changing the receptor copy numbers, see SI. Since this is observed experimentally (Yu and Wu, 2016), we predict that the receptor copy numbers are adjusted such that the exci- tations of all glomeruli are similar when averaged over (cid:31)an(cid:30) = 0.5 (cid:31)an(cid:30) = 0.2 (cid:31)an(cid:30) = 0.1 0.2 0.4 0.6 0.8 1 30 20 10 C Infor.I[bits] 0 0 (cid:31)an(cid:30) = 0.5 (cid:31)an(cid:30) = 0.2 (cid:31)an(cid:30) = 0.1 0.5 1 0 1.5 2 Sensitivity ξ1 of receptor 1 Sensitivity variation var(ξn)/(cid:31)ξn(cid:30)2 1 0.5 0 −0.5 −1 I−I0[bits] B ξ1 = 1 ξ1 = 2 5 10 15 20 25 30 0 0.03 Channel n Frequency 60 40 20 0 60 40 20 0 A en en natural odors. Alternatively, variations in the receptor sensitivities could be balanced by more complex inhibi- tion mechanism. For instance, experiments show that different projection neurons have different susceptibili- ties to inhibition (Hong and Wilson, 2015). Here, the experimentally observed turnover of mitral cells and in- terneurons (Lazarini and Lledo, 2011) could adjust the inhibition mechanism locally, which could optimize the olfactory system for a given environment (Mouret et al., 2009). Such adaptation of the inhibition mechanism to the current stimulus statistics and more complex mod- els where the behavioral state of an animal could influ- ence the olfactory bulb by top-down modulation (Wilson, 2013) will be interesting to explorer in the future. Our simplified model neglects many details of the olfac- tory system (Silva Teixeira, Cerqueira, and Silva Ferreira, 2016). For instance, we do not consider the dynamics of inhalation and the odor absorption in the mucus (Pelosi, 2001; Schoenfeld and Cleland, 2005). Instead, we here directly parameterize the ligand distribution at the ol- factory receptors, where we for simplicity neglect corre- lations between ligands. It would be interesting to extend the model for more complex stimuli and study how the system decorrelates the input, identifies a target odor in a background, and separates multiple odors from each other. This likely involves many steps (Cleland et al., 2011) and cannot be done perfectly with a single normal- ization step and non-linear gain function. For instance, it might be important to apply gain functions at the level of receptors and the glomeruli to model finite sensitivity and saturation effects. Additionally, it has been shown that there is additional cross-talk on the level of recep- tors (Ukhanov et al., 2010) and glomeruli (Aungst et al., 2003; Silbering and Galizia, 2007), which could support decorrelation. Generally, such cross-talk and the inhibi- tion that we discussed here will be non-linear (Wilson, 2011). This could for instance be modeled by a divisive normalization model that has been proposed for olfac- tion (Olsen, Bhandawat, and Wilson, 2010). It is also likely that the inhibition of the projection neurons is not driven by a single global variable. If glomeruli position- ing carried some meaning (Murthy, 2011), local inhibi- tion could help separating similar odors by enhancing the contrast (Leon and Johnson, 2003). The discrimi- nation of similar odors could also be improved if projec- tion neurons had a larger output range, increasing the in- formation capacity per channel. Finally, we completely neglected the temporal dynamics of the olfactory sys- tem, which play an important role for the adaptation be- tween sniffs (Zufall and Leinders-Zufall, 2000) and might also influence odor perception within a single sniff (Blau- velt et al., 2013; Sirotin, Shusterman, and Rinberg, 2015; Uchida, Poo, and Haddad, 2014). 9 ACKNOWLEDGMENTS I thank Michael P. Brenner, Venkatesh N. Murthy, Mikhail Tikhonov and Christoph A. Weber for helpful discussions and a critical reading of the manuscript. This research was funded by the Simons Foundation and the German Science Foundation through ZW 222/1-1. REFERENCES Asahina, K., Louis, M., Piccinotti, S., and Vosshall, L. B., J Biol 8, 9 (2009). Aungst, J. L., Heyward, P. M., Puche, A. C., Karnup, S. V., Hayar, A., Szabo, G., and Shipley, M. T., Nature 426, 623 (2003). Banerjee, A., Marbach, F., Anselmi, F., Koh, M. S., Davis, M. B., Garcia da Silva, P., Delevich, K., Oyibo, H. K., Gupta, P., Li, B., and Albeanu, D. F., Neuron 87, 193 (2015). Barlow, H., Network 12, 241 (2001). Barlow, H. B., in Sensory Communication, edited by W. Rosenblith (MIT press, 1961) pp. 217–234. Berck, M. E., Khandelwal, A., Claus, L., Hernandez-Nunez, L., Si, G., Tabone, C. J., Li, F., Truman, J. W., Fetter, R. D., Louis, M., Samuel, A. D., and Cardona, A., Elife 5 (2016), 10.7554/eLife.14859. Bhandawat, V., Olsen, S. R., Gouwens, N. W., Schlief, M. L., and Wilson, R. I., Nat. Neurosci. 10, 1474 (2007). Blauvelt, D. G., Sato, T. F., Wienisch, M., V. N., Frontiers in neural circuits 7 (2013). and Murthy, Bushdid, C., Magnasco, M., Vosshall, L., and Keller, A., Science 343, 1370 (2014). Cain, W. S., Science 195, 796 (1977). Carandini, M. and Heeger, D. J., Nat Rev Neurosci 13, 51 (2012). Cleland, T. A., Trends. Neurosci. 33, 130 (2010). Cleland, T. A., Chen, S.-Y. T., Hozer, K. W., Ukatu, H. N., Wong, K. J., and Zheng, F., Front Neuroeng 4, 21 (2011). Cleland, T. A., Johnson, B. A., Leon, M., and Linster, C., Proc. Natl. Acad. Sci. USA 104, 1953 (2007). Cleland, T. A. and Sethupathy, P., BMC Neurosci 7, 7 (2006). Davison, I. G. and Katz, L. C., J Neurosci 27, 2091 (2007). Demb, J. B. and Singer, J. H., Annual Review of Vision Sci- ence 1, 263 (2015). Getz, W. M. and Lutz, A., Chem. Senses 24, 351 (1999). Goyert, H. F., Frank, M. E., Gent, J. F., and Hettinger, T. P., Brain Res Bull 72, 1 (2007). Gupta, P., Albeanu, D. F., and Bhalla, U. S., Nat. Neurosci. 18, 272 (2015). Haddad, R., Weiss, T., Khan, R., Nadler, B., Mandairon, N., Bensafi, M., Schneidman, E., and Sobel, N., J Neurosci 30, 9017 (2010). Hong, E. J. and Wilson, R. I., Neuron 85, 573 (2015). Hopfield, J., Proc. Natl. Acad. Sci. USA 96, 12506 (1999). Isaacson, J. S. and Scanziani, M., Neuron 72, 231 (2011). Jefferis, G. S., Marin, E. C., Stocker, R. F., and Luo, L., Nature 414, 204 (2001). Jinks, A. and Laing, D. G., Perception 28, 395 (1999). Jinks, A. and Laing, D. G., Physiol Behav 72, 51 (2001). Kaupp, U. B., Nat Rev Neurosci 11, 188 (2010). Koulakov, A., Gelperin, A., and Rinberg, D., J. Neurophysiol. 98, 3134 (2007). Laurent, G., Science 286, 723 (1999). Lazarini, F. and Lledo, P.-M., Trends Neurosci 34, 20 (2011). Leon, M. and Johnson, B. A., Brain Res. Rev. 42, 23 (2003). Li, Z., Biol Cybern 62, 349 (1990). Li, Z., in Models of neural networks (Springer, 1994) Chap. 6, pp. 221–251. Lin, D. Y., Zhang, S.-Z., Block, E., and Katz, L. C., Nature 434, 470 (2005). Stevens, C. F., Proc. Natl. Acad. Sci. USA 112, 9460 (2015). Stevens, C. F., Proc. Natl. Acad. Sci. USA (2016), 10.1073/pnas.1606339113. Stitzel, S. E., Aernecke, M. J., and Walt, D. R., Annu. Rev. Biomed. Eng. 13, 1 (2011). Su, C.-Y., Menuz, K., and Carlson, J. R., Cell 139, 45 (2009). Tabor, R., Yaksi, E., Weislogel, J.-M., and Friedrich, R. W., J. Neurosci. 24, 6611 (2004). Linster, C. and Hasselmo, M., Behavioural brain research 84, Tan, J., Savigner, A., Ma, M., and Luo, M., Neuron 65, 912 117 (1997). (2010). Lowe, G. and Gold, G. H., Proc. Natl. Acad. Sci. USA 92, Tikhonov, M., Little, S. C., and Gregor, T., R Soc Open Sci 10 7864 (1995). 2, 150486 (2015). Mainland, J. D., Lundstrom, J. N., Reisert, J., and Lowe, Tkacik, G. and G., Trends Neurosci. 37, 443 (2014). Malnic, B., Hirono, J., Sato, T., and Buck, L. B., Cell 96, 713 (1999). Mouret, A., Lepousez, G., Gras, J., Gabellec, M.-M., and Lledo, P.-M., J Neurosci 29, 12302 (2009). Murthy, V. N., Annu. Rev. Neurosci. 34, 233 (2011). Niimura, Y., Curr Genomics 13, 103 (2012). Nikolova, N. and Jaworska, J., QSAR & Combinatorial Sci- ence 22, 1006 (2003). Olsen, S. R., Bhandawat, V., and Wilson, R. I., Neuron 66, 287 (2010). Olshausen, B. A. and Field, D. J., Curr Opin Neurobiol 14, 481 (2004). Pelosi, P., Cellular and Molecular Life Sciences CMLS 58, 503 (2001). Perez-Orive, J., Mazor, O., Turner, G. C., Cassenaer, S., Wil- son, R. I., and Laurent, G., Science 297, 359 (2002). Rinberg, D., Koulakov, A., and Gelperin, A., J Neurosci 26, 8857 (2006). Roland, B., Jordan, R., Sosulski, D. L., Diodato, A., Fuku- naga, I., Wickersham, I., Franks, K. M., Schaefer, A. T., and Fleischmann, A., Elife 5 (2016), 10.7554/eLife.16335. Ruderman, and Bialek,, Phys. Rev. Lett. 73, 814 (1994). Sachse, S. and Galizia, C. G., J Neurophysiol 87, 1106 (2002). Schoenfeld, T. A. and Cleland, T. A., Trends in neurosciences 28, 620 (2005). Silbering, A. F. and Galizia, C. G., J. Neurosci. 27, 11966 (2007). Silva Teixeira, C. S., Cerqueira, N. M. F. S. A., and Silva Fer- reira, A. C., Chem Senses 41, 105 (2016). Sirotin, Y. B., Shusterman, R., and Rinberg, D., eNeuro 2 (2015), 10.1523/ENEURO.0083-15.2015. Soucy, E. R., Albeanu, D. F., Fantana, A. L., Murthy, V. N., and Meister, M., Nat. Neurosci. 12, 210 (2009). Bialek, W., Physics Annual 89 7, Review (2016), Condensed Matter of http://dx.doi.org/10.1146/annurev-conmatphys-031214- 014803. Uchida, N. and Mainen, Z. F., Front Syst Neurosci 1, 3 (2007). Uchida, N., Poo, C., and Haddad, R., Annu Rev Neurosci 37, 363 (2014). Ukhanov, K., Corey, E. A., Brunert, D., Klasen, K., and Ache, B. W., J Neurophysiol 103, 1114 (2010). Verbeurgt, C., Wilkin, F., Tarabichi, M., Gregoire, F., Du- and Chatelain, P., PLOS ONE 9, e96333 mont, J. E., (2014). Weiss, T., Snitz, K., Yablonka, A., Khan, R. M., Gafsou, D., Schneidman, E., and Sobel, N., Proc. Natl. Acad. Sci. USA 109, 19959 (2012). Wilson, R. I., Curr. Opin. Neurobiol. 21, 254 (2011). Wilson, R. I., Annu. Rev. Neurosci. 36, 217 (2013). Wright, G. A. and Thomson, M. G., in Integrative Plant Bio- chemistry, Recent Advances in Phytochemistry, Vol. 39, edited by J. Romeo (Elsevier, 2005) Chap. 8, pp. 191–226. Yokoi, M., Mori, K., and Nakanishi, S., Proc. Natl. Acad. Sci. USA 92, 3371 (1995). Yu, C. R. and Wu, Y., Exp Neurol (2016), 10.1016/j.expneurol.2016.06.001. Yu, Y., Claire, A., Ni, M. J., Adipietro, K. A., Golebiowski, J., Matsunami, H., and Ma, M., Proc. Natl. Acad. Sci. USA 112, 14966 (2015). Zhang, D., Li, Y., and Wu, S., Comput Math Methods Med 2013, 507143 (2013). Zufall, F. and Leinders-Zufall, T., Chem. Senses 25, 473 (2000). Zwicker, D., Murugan, A., and Brenner, M. P., Proc. Natl. Acad. Sci. USA 113, 5570 (2016). Supporting Information: Normalized neural representations of natural odors David Zwicker1, 2, ∗ 1School of Engineering and Applied Sciences, Harvard University, Cambridge, MA 02138, USA 2Kavli Institute for Bionano Science and Technology, Harvard University, Cambridge, MA 02138, USA (Dated: July 3, 2018) (a variable) number of log-normally distributed random variables, its distribution can be approximated by an- other log-normal distribution (Wu, Mehta, and Zhang, 2005), which we parameterize by its mean µtot and vari- ance σ2 tot. We consider the simple approximation where these parameters are directly given by Eq. S2 (Fenton, 1960). This choice approximates the tail of the distri- bution well, but leads to errors in the vicinity of the mean (Wu, Mehta, and Zhang, 2005). Since both ctot and ci are log-normally distributed when ligand i is present in an odor (ci > 0), ci is also log-normally distributed in this case and hciici>0 = var(ci)ci>0 = χ µi µtot µ2 i χ2 µ2 totµ−2 tot (cid:18) σ2 i µ2 i where χ = 1 + σ2 1 − pi, the statistics of ci read χ + χ − 1(cid:19) , (S4a) (S4b) tot. Since ci = 0 with probability hcii = var(ci) = χ piµi µtot piµ2 µ2 i χ2 tot (cid:18) σ2 i µ2 i χ + χ − pi(cid:19) . (S5a) (S5b) Note that the covariance cov(ci, cj) does not vanish since the ci are not independent. In particular, var(Pi ci) = 0, since Pi ci = 1 by definition. This condition is only of the normalized excitations en = ¯S−1Pi Snici read consistent with Eq. S5a if χ ≈ 1, which implies that ctot must not vary much, σtot µtot (cid:28) 1. Using χ = 1, the statistics (S6a) heni = 1 var(en) = var(Sni) ¯ S2 DXi c2 iE , (S6b) where hPi c2 the statistics given in Eq. S5. ii = hcii2 + var(ci) and In the simple case where all ligands are drawn from the ii with hc2 ii ≈Pihc2 same distribution (pi = p, µi = µ, σi = σ), we obtain , (S7) , (S8) (cid:18)1 + σ2 µ2(cid:19) var(Sni) ¯ S2 1 s var(en) ≈ 1 − p + σ2 µ2 sNL , var(ci) ≈ L ( σ2 µ2 + 1), such that 1 N hcii ≈ ii ≈ 1 sNL and hc2 which is equivalent to Eq. 5 in the main text. CONTENTS S1. Statistics of normalized concentrations and excitations S2. Numerical simulations S3. Approximate channel activity S4. Odor discriminability S5. Receptor binding model 1 2 2 2 3 S1. STATISTICS OF NORMALIZED CONCENTRATIONS AND EXCITATIONS Let pi be the probability that ligand i is present in an odor. If it is present, its concentration ci is drawn from a log-normal distribution with mean µi and standard devi- ation σi, while ci = 0 if the ligand is not present. Hence, hcii = piµi var(ci) = (pi − p2 i )µ2 i + piσ2 i , (S1a) (S1b) while the covariances cov(ci, cj) = hcicji−hciihcji vanish for i 6= j since the ligands are independent. The statistics of the total concentration ctot =Pi ci read NLXi=1 hctoti = NLXi=1 var(ctot) = hcii and The excitations en are given by en =Pi Snici, where the sensitivities Sni are log-normally distributed with mean hSnii = ¯S and variance var(Sni) = ¯S2(eλ2 − 1). Hence, (S3a) heni = ¯Shctoti var(ci) . (S2) var(en) = ¯S2 var(ctot) + var(Sni) NLXi=1 hc2 ii , (S3b) i + σ2 ii = pi(µ2 where hc2 i ) and cov(en, em) = 0 for n 6= m. We next determine the statistics of the normalized con- centrations ci = ci/ctot. For simplicity, we consider large odors, Pi pi (cid:29) 1, where ctot can be considered as an independent random variable. Since ctot is the sum of ∗ [email protected]; http://www.david-zwicker.de arXiv:1608.01179v1 [q-bio.NC] 3 Aug 2016 associated probability distribution function reads 2 (S10) (S11) # 2S2 n exp"−(cid:0)Mn − ln(en)(cid:1)2 (cid:21) . erfc(cid:20) Mn − ln(en) √2Sn and the cumulative distribution function is 1 2 F (en) = f (en) = 1 √2πSnen The parameters Mn and Sn can be determined from the mean and variance var(en) = e2Mn+S2 S2 n 2 (cid:19) heni = exp(cid:18)Mn + n − 1(cid:17) . n(cid:16)eS2 Sn =p2ζ , and Mn = lnheni − ζ Solving these equations for Mn and Sn, we obtain (S12a) (S12b) (S13) hani ≈ where ζ = 1 2 ln(1 + var(en)heni−2). Eq. 6 of the main text follows from this and Eq. 5. For small hani we have 2pζ/π 4ζ /(x√π), valid for which follows from erfc(x) ≈ e−x2 x (cid:29) 1. For small ζ, we obtain the approximate scal- ing lnhani ∼ −(ln α)2/(4ζ), where ζ ∼ s−1 for s (cid:29) 1. exp(cid:20)− (ln(α) + ζ)2 (cid:21) , ln(α) + ζ (S14) S4. ODOR DISCRIMINABILITY We quantify the discriminability of two odors by the Hamming distance d of their respective representations a for several different cases: a. Uncorrelated odors The expected distance hdi be- tween the activity patterns a(1) and a(2) of two inde- pendent odors is n i + ha(2) hdi = NR(cid:16)ha(1) n i and ha(2) where ha(1) n i denote the expected activities of the two odors, averaged over sensitivity matrices, and we neglect correlations cov(an, am) for simplicity. n i(cid:17) , n iha(2) n i − 2ha(1) (S15) FIG. S1 Influence of the width λ of the sensitivity distribu- tion on the statistics of the odor representations. (A) Ex- pected channel activity hani as a function of λ for several inhibition strengths α. Intermediated values, λ ≈ 1, lead to larger activities. (B) Mean Pearson correlation coefficient hρi calculated from an ensemble average of Eq. S9 as a function of λ for several α. For small λ, hani was too small to esti- mate ρ reliably. (A–B) Results are shown for small ( σ µ = 1, solid lines) and large ( σ µ = 10, dashed lines) concentration variability. Remaining parameters are NR = 32, NL = 256, and pi = 0.1. S2. NUMERICAL SIMULATIONS We numerically calculated ensemble averages over odors c and sensitivity matrices Sni. Here, we first choose Sni by drawing all entries independently from a log-normal distribution with mean ¯S = 1 and vari- ance var(Sni) = eλ2 − 1. We then draw an odor c using the following procedure: First, we determine which of the NL ligands are present according to their probabili- ties pi. Second, we draw the concentrations ci for each ligand i that is present from a log-normal distribution with mean µi and standard deviation σi. We then use Eqs. 1–3 given in the main text to map the odor c to a binary activity vector a, from which we can for in- stance calculate the number of active channels. We ob- tain ensemble averages of such quantities by repeating these steps 105 times. This allows us to calculate the mean activities hani, the covariances cov(an, am), and the Pearson correlation coefficient ρ, which is defined as ρ = NR 1 2 − NR Xn6=m cov(an, am) (cid:2)var(an) var(am)(cid:3) 1 2 . (S9) Fig. S1 shows these quantities as a function of the width λ of the sensitivity distribution. We also estimate P (a) from an ensemble average to calculate the information I from its definition given in Eq. 4 in the main text. S3. APPROXIMATE CHANNEL ACTIVITY We estimate the expected activity hani by the proba- bility that the normalized excitations en exceed the ex- pected normalized threshold α. Since both the sensitivi- ties Sni and the normalized concentrations ci are approx- imately log-normally distributed, en can also be approx- imated by a log-normal distribution (Fenton, 1960). The b. Adding target to background We calculate the ex- pected change hdi of the representation when a target odor ct is added to a background odor cb. Because the odor concentrations are specified, we consider the actual excitations en instead of the normalized quantities en. Taking an ensemble average over sensitivity matrices, the excitations associated with the two odors are charac- terized by probability distribution functions f t E(et) and 1 % 0 % α = 1.5 α = 2 α = 3 0 3 Sensitivity distribution width λ 2 1 −1 % −2 % Correlation(cid:31)ρ(cid:30) B 10−4 0.1 10 Sensitivity distribution width λ 1 1 0.1 0.01 α = 1.5 α = 2 α = 3 10−3 Activity(cid:31)an(cid:30) A f b E(eb) for the target and the background, respectively. We here consider log-normally distributed en, which are parameterized by their mean and variance, heni = ¯S NLXi=1 ci var(en) = var(Sni) c2 i , (S16) NLXi=1 where var(Sni) = ¯S2(eλ2 − 1). tion heκi =R z f κ When the target is added to the background, the expected threshold hγi increases from γb = αhebi to γs = α(hebi + heti), where heκi denotes the mean excita- E(z) dz for κ = t, b. This increase in the threshold can deactivate a channel if it was previously active, i. e. if its excitation was larger than the thresh- old associated with the background, eb > γb. For such eb, the probability that the receptor gets deactivated by adding the target is P (eb + et < γseb). Integrating over all possible eb, we thus get the probability poff that a channel becomes inactive, poff =Z ∞ =Z ∞ γb γb P(cid:0)eb + et < γs(cid:12)(cid:12) eb(cid:1) f b E(cid:0)eb(cid:1) deb , E(cid:0)γs − eb(cid:1)f b F t E(cid:0)eb(cid:1) deb (S17) E(et) is the cumulative distribution function as- where F t sociated with f t E(et). Conversely, a channel becomes ac- tive when the additional excitation by the target odor brings it above the threshold γs. The associated proba- bility pon reads pon =Z γb 0 (cid:2)1 − F t E(cid:0)γs − eb(cid:1)(cid:3) f b E(cid:0)eb(cid:1) deb . Taken together, the expected number hdi of channels that change their state reads hdi = NR ·(cid:0)pon + poff(cid:1) . There are three simple limits that we can solve analyt- If there is no target, heti = 0, the activation ically: pattern does not change and we have hdi = 0. In the opposing limit of a dominant target, heti → ∞, the acti- vation patterns are independent and we recover the dis- tance hdimax for uncorrelated odors, which is given by Eq. S15. Lastly, in the case where the target and the background are identically distributed, hebi = heti and var(eb) = var(et), we have hdi = 1 2hdimax. (S18) (S19) 3 cause a baseline excitation eb, which is distributed ac- E(eb). A channel is inactive for an odor with cording to f b E(hγi − eb), where F d probability F d E(ed) is the cumula- tive distribution function of the excitation caused by the sd = s − sb different ligands. Hence, p = 2Z hγi 0 F d E(z)(cid:2)1 − F d E(z)(cid:3)f b E(eb)deb , (S20) where z = hγi− eb. Note that the upper bound of the in- tegral is hγi since channels will be active for both odors if eb ≥ hγi. The associated Hamming distance hdi between the two odors is then given by hdi = pNR. S5. RECEPTOR BINDING MODEL We consider a simple model where receptors Rn get activated when they bind ligands Li. This binding is described by the chemical reaction Rn + Li (cid:10) RnLi, where RnLi is the receptor-ligand complex. In equilib- rium, the concentrations denoted by square brackets obey [RnLi] = Kni·[Rn][Li], where Kni is the binding constant of the reaction. Hence, crec n Knici , (S21) [RnLi] = 1 +Pi Knici n = [Rn]+Pi[RnLi] denotes where we consider the case where multiple ligands com- pete for the same receptor. Here, ci = [Li] is the concen- tration of free ligands and crec the fixed concentration of receptors, which is related to the copy number of receptors of type n. We consider a simple receptor model where the excitation is propor- tional to the concentration of the bound ligands, such that the excitation accumulated in glomerulus n reads en = βn N rec n crec n NLXi=1 [RnLi] = βnN rec . (S22) n Pi Knici 1 +Pi Knici n Here, N rec is the copy number of receptors of type n and βn characterizes their excitability, which could for in- stance be modified by point mutations (Yu et al., 2015). Defining Sni = βnN rec n Kni, we recover Eq. 1 of the main text in the limit of small concentrations,Pi Knici (cid:28) 1. The sensitivities are thus proportional to the copy num- ber N rec and the biochemical details encoded in βnKni. n REFERENCES c. Discriminating two odors of equal size We consider the simple case of two odors that each contain s ligands at equal concentration, sharing sb of them, such that the ex- pected threshold hγi is the same for both odors. Similar to the derivation above, we here calculate the probabil- ity p that a channel is active for one odor, but not for the other. The sb ligands that are present in both odors Fenton, L. F., Communications Systems, IRE Transactions on 8, 57 (1960). Wu, J., Mehta, N. B., and Zhang, J., in GLOBECOM '05. IEEE Global Telecommunications Conference, 2005., Vol. 6 (2005) pp. 3413–3417. Yu, Y., Claire, A., Ni, M. J., Adipietro, K. A., Golebiowski, J., Matsunami, H., and Ma, M., Proc. Natl. Acad. Sci. USA 112, 14966 (2015).
1702.03579
1
1702
2017-02-12T21:18:10
Autonomous line follower robot controlled by cell culture
[ "q-bio.NC" ]
Neuro-electronic hybrid promises to bring up a model architecture for computing. Such computing architecture could help to bring the power of biological connection and electronic circuits together for better computing paradigm. Such paradigms for solving real world tasks with higher accuracy is on demand now. A robot as a autonomous system is modeled here to navigate following a particular line. Sensory inputs from robot is directed as input to the cell culture in response to which motor commands are generated from the culture.
q-bio.NC
q-bio
Autonomous line follower robot controlled by cell culture Sayan Biswas Department of Electrical Engineering Jadavpur University, Kolkata, India Email: [email protected] 7 1 0 2 b e F 2 1 ] . C N o i b - q [ 1 v 9 7 5 3 0 . 2 0 7 1 : v i X r a Abstract-Neuro-electronic hybrid promises to bring up a model architecture for computing. Such computing architecture could help to bring the power of biological connection and electronic circuits together for better computing paradigm. Such paradigms for solving real world tasks with higher accuracy is on demand now. A robot as a autonomous system is modeled here to navigate following a particular line. Sensory inputs from robot is directed as input to the cell culture in response to which motor commands are generated from the culture. Index Terms-Neuro-electronic hybrid, biological connection, line follower, sensory electronic circuits, autonomous system, inputs, motor commands I. INTRODUCTION Computation carried by brain are fast and they are able to take rapid decisions. At times brain can take really fast decision at near zero reflex time. Although all kind of decision making does not fall into this category, but obviously there are instances where one can take decision in near zero reflex time. In a visual search task to discriminate between a car image and animal image, near zero reflex time is expected. Decision making regarding colours whether is black or whites takes almost zero reflex time. Such computation are carried by brain real fast. Such abilities of the brain have inspired researchers. Many strategies have been developed inspired from such abilities of brain. Some of those strategies are neuromorphic devices [1] [2] [3], artificial neural network [4], fuzzy systems [5] [6]. An ongoing research is being carried out for making hardwares suitable to mimic neuronal systems [1]. There are challenges including versatility connectivity which limits the usage of cul- tures like brain. Neuro electronic hybrid systems can prove to be a possible computing architecture. On connecting cultures of neurons properly with the real world such abilities of the brain can be used to solve real world problems. This might prove to be a essential platform to use biological network for computing. Neurons are basic computation and functional unit of brain which are versatile in terms of behavior. The neurons in the brain form inter connected set of networks which are dynamic in nature. Such dynamic nature of neuron in network gives them learning abilities and makes them powerful enough to serve varied types of function. Hence making the brain a robust computational unit of brain. A hybrid system attempts to solve real world problems by implementing this abilities of brain. MEA dish proves advantageous and promising for under- standing network activity and properties of neuronal network. Robotic control through such culture [7] [8] and training [9] of neuronal network is an ongoing work at various labs. There are several challenges required to be resolved for a practical and real world application of such systems [10]. Decoding output from an input to a network is open problem. To understand such decoding pattern a deeper understanding of dynamic network system [11] [12] [13] [14] [15] is essential. The description here deals with a framework for using cell cultures as a information processing tool, and to implement the ability for the purpose of classification of sensory inputs received and accordingly take decision of generating correct motor commands. The framework shall elaborately explain a proposal for implementation in real life. II. LINE FOLLOWER A line follower robot is an autonomous body expected to navigate in a network by following a specific line. The track on which the robot is expected to navigate is coloured black, and the background is white (figure 2). Such devices are controlled by a person using a visual feedback from the scene of the track which is used to generate respective motor commands on the controller to control the robot. This involves multi sensory perception to compute the action required to be taken from the visual input and act accordingly by sending motor commands to the fingers to produce the desired effect on the robot navigating on the track. As this requires intervention in control from a human this can't be refereed to as a automated system. The process is as shown in the figure 1. Fig. 1: Robot Navigation by Human Several computer vision [16] [17] along with machine lean- ing [18] paradigm has been successful to implement the same Fig. 2: TRACK by making silicon chips to perform the computation that are to be performed by the brain. There are cameras on the robot which are performing the role of human eyes. The images from the camera gives some features regarding the path on which the robot is navigating. The direction on which the robot must move is classified based on the features extraction method and using a machine learning classifier. Once this is known, motor commands to make the robot move in desired direction is sent to the wheels. Hence the silicon chips and camera do the work of eye and multi sensory perception [19], of relating visual feedback to generate motor commands. Using computer vision techniques for extracting visual features followed by machine learning classifier to generate motor commands is shown in the figure 3. As specified the aim is to implement cell culture to make the controlling of robot automated. III. CELL CULTURES AS CONTROLLING UNIT Multi electrode array - MEA [20] recording gives a great opportunity to study network topology as it provides popu- lation recording. It provides data from multiple electrodes or channels which is thus effective for understanding network events. A MEA has culture on its surface from which electrical activity is recorded. A framework is modeled for using the electrical activity recording to control a robot to navigate following a line. Involvement of cell as a contriving unit will be done by following: • By photo receptors cells • By neural cells. These two shall be dealt in following section. IV. PHOTO RECEPTOR CELLS Photo receptor cells are specialized cells in retina which performs photo transduction. Visual photo transduction is the method of converting light into electrical signals. Photo receptors are of great importance as they convert light into sig- nals which stimulate biological processes. These cells absorb photons that causes a change in the cell's membrane potential. The model proposes usage of such cell culture on MEA dish for controlling a line follower. The robot would be a simple system equipped with a camera to serve the purpose of visual feedback. As the camera takes up the current position of the robot, it is to be projected on the MEA dish containing photo receptor cell cultures. The usage of camera serves the purpose of eye. The projection of the visual field is shown in the figure 5. The track (figure 2) if followed carefully one can find there are three possible different visual field as shown in figure 6. The photo recep- tors respond by different electrical activity to different light intensities. Hence the projection of different visual field on the culture is expected to generate different electrical activity MEA dish. Hence to achieve control using photo receptor cell culture [21] it requires a mapping of electrical activity to visual field projected. This mapping would help to understand the visual field or current location of robot with the electrical activity recorded which would in turn help in generating motor commands. After the culture is done three of this possible visual field would be projected one by one, and corresponding electrical activity generated would be recorded. This will help Fig. 3: Robot Navigation by Machine Learning classifier TABLE I: Motor Commands as per the Visual Field Visual Field Visual Field A Visual Field B Visual Field C Motor Commands Cover offset and take 90 ◦ left Cover offset and take 90 ◦ right Go straight in achieving the required mapping of visual field and electrical activity mapping. Once the mapping is achieved the robot equipped with camera is left to be navigate on path. It would take visual field project it on the culture followed by which electrical activity shall be recorded based on which proper motor command would be generated and sent to the robot. The methodology is as follows: • Visual fields are projected on to the cell culture and mapping of electrical activity and visual field is obtained • During test, the images are projected which will generate motor commands as per obtained mapping As per positioning of the camera in the experiment an offset has to determined. The offset would prevent the robot to take desired motor action as soon as visual field is detected. Offset is shown in figure 4. If the robot would take say left turn on attending visual field A or electrical activity corresponding to visual field A, it would go off track out of line. This demands necessity of having a pre determined offset. Hence motor commands required to be generated as per the visual field is given in table I. The procedure is illustrated in figure 7. V. NEURON CELL Neuron cultures obtained generally from rat brain are cul- tured on MEA dish can be implemented as a control unit for making autonomous navigator. Framework that will be discussed here was proposed earlier [10] for a autonomous Fig. 4: Offset Fig. 5: Projection of the visual field captured through camera on top of the MEA dish containing photoreceptor cells (a) Visual field - A (b) Visual field - B (c) Visual field - C Fig. 6: Visual field projected on MEA dish Fig. 7: Automated robot navigation by photo receptor cell (a) Sensor position - A (c) Sensor position - C Fig. 8: Sensor position of robot as it navigates. Blue circles denotes location of sensor. (b) Sensor position - B Fig. 9: Automated robot navigation by neural cell TABLE II: Motor Commands as per the Sensor position Sensor Position Sensor Position A Sensor Position B Sensor Position C Motor Commands Take 90 ◦ left and a little forward drift Take 90 ◦ right and a little forward drift Go straight Fig. 10: Result of left turn from sensor position A - figure 8a. Blue circles denotes location of sensor. system of navigating by avoiding obstacles. The proposed framework had a accuracy about 98%. It would be discussed how already existing framework could be utilized for modeling a line follower system. The robot which shall navigate must have two infrared - IR sensors positioned pointing downward. As the robot navigates there could the only three possible states of view as shown in figure 8. A IR sensor contains a emitter and detector. The emitter would emit infrared waves and purpose of detector would be to detect presence of any IR wave. As white reflects light and black absorbs it, same would be true with infrared wave. Hence if IR sensor detect (1) IR waves that means the colour is white and if it does not detect (0) the colour must be black. Hence {1,1} {0,1} {1,0} corresponds to go straight, go left, go right respectively (figure 8 gives a pictorial representation). Hence knowing the output of IR sensor could predict the desired output motor command. Activity of MEA dish is used to understand the output of IR sensor. It was found that probability of spike occurrence at some particular electrode in response to stimulus sequence at some other elec- trode depends on electrode stimulation the order of stimulation and timing between the pulses [22]. A network of neuron is well able to distinguish between various inputs based on electrodes stimulated, the stimulating order and timing [10]. This ability of neuronal network was used to encode inputs to a culture and decode output responses from the culture. From electrodes that excited the network to highest level stimulating electrode is chosen. Stimulating and recording electrodes are so chosen that by input pattern could be predicted from activity at recording electrodes [10]. Hence the input pattern of {1,1} {0,1} {1,0} could be predicted from activity patter of neuron cultures. As this prediction could be made foreseen motor commands could be generated making it possible for the robot navigate following a line. Suppose the sensor position is as in figure 8a. If then robot takes a 90 ◦ left turn it results to a position described in figure 10 which is a undesired position. Similarly the system would attain a undesired position if robot takes a 90 ◦ right turn from position C (figure 8c). This is undesired as one of the sensor is lying on white region where as the other is on black region. This issue could be resolved by generating a motor command to take a 90 ◦ turn on the desired direction followed by a automated little forward push so that both sensor come into the white region. Hence automated navigator using neuron culture is developed. The corresponding motor commands are shown in table II. The procedure is illustrated in figure 9. VI. DISCUSSION AND CONCLUSIONS This was a implementation model using two type of cell culture in performing real world task of navigating following a line. It was shown how could the system be made autonomous by using two different type of cell cultures. The techniques for usage of neuron culture have been implemented practically and was found to have a accuracy of 98% [10]. It is expected that if the same methodology is extended in autonomous line follower system, as proposed, it would have a good accuracy too. The strong idea in this approach is using cell cultures as information processing tool, and to implement the ability for the purpose of classification of sensory inputs received and accordingly take decision of generating correct motor commands. Driving can be looked as a process of locomotion by obstacle avoidance and path following. A amalgamation of object avoiding paradigm [10] and line follower techniques, as suggested in thid work, could pave way for technological ad- vancement of building new efficient and accurate autonomous driving systems. VII. ACKNOWLEDGMENT Author would like to thank Shefali, Department of Food Technology and Biochemical Engineering, Jadavpur Univer- sity, Kolkata for the valuable comments in improving the manuscript. Author would also like to thank Department of Electrical Engineering, Jadavpur University, Kolkata. REFERENCES [1] M. Suri, O. Bichler, D. Querlioz, O. Cueto, L. Perniola, V. Sousa, D. Vuillaume, C. Gamrat, and B. DeSalvo, "Phase change memory as synapse for ultra-dense neuromorphic systems: Application to complex visual pattern extraction," in Electron Devices Meeting (IEDM), 2011 IEEE International. IEEE, 2011, pp. 4–4. [2] S. Yu, Y. Wu, R. Jeyasingh, D. Kuzum, and H.-S. P. Wong, "An electronic synapse device based on metal oxide resistive switching memory for neuromorphic computation," IEEE Transactions on Electron Devices, vol. 58, no. 8, pp. 2729–2737, 2011. [3] G. Indiveri, "A neuromorphic vlsi device for implementing 2d selective attention systems," IEEE Transactions on Neural Networks, vol. 12, no. 6, pp. 1455–1463, 2001. [4] J. E. Dayhoff and J. M. DeLeo, "Artificial neural networks," Cancer, vol. 91, no. S8, pp. 1615–1635, 2001. [5] D. Nauck, F. Klawonn, and R. Kruse, Foundations of neuro-fuzzy systems. John Wiley & Sons, Inc., 1997. [6] C.-T. Lin and C. Lee, Neural fuzzy systems: a neuro-fuzzy synergism to intelligent systems. Prentice-Hall, Inc., 1996. [7] D. J. Bakkum, Z. C. Chao, and S. M. Potter, "Spatio-temporal electrical stimuli shape behavior of an embodied cortical network in a goal- directed learning task," Journal of neural engineering, vol. 5, no. 3, p. 310, 2008. [8] A. Novellino, P. D'angelo, L. Cozzi, M. Chiappalone, V. Sanguineti, and S. Martinoia, "Connecting neurons to a mobile robot: an in vitro bidirec- tional neural interface," Computational Intelligence and Neuroscience, vol. 2007, 2007. [9] M. E. Ruaro, P. Bonifazi, and V. Torre, "Toward the neurocomputer: image processing and pattern recognition with neuronal cultures," IEEE Transactions on Biomedical Engineering, vol. 52, no. 3, pp. 371–383, 2005. [10] J. B. George, G. M. Abraham, B. Amrutur, and S. K. Sikdar, "Robot navigation using neuro-electronic hybrid systems," in 2015 28th Inter- national Conference on VLSI Design. IEEE, 2015, pp. 93–98. [11] S. Biswas, "Proposal for ranking nodes of neural network using activity index," in Research in Computational Intelligence and Communication Networks (ICRCICN), 2016 Second International Conference on. IEEE, 2016, pp. 224–228. [12] S. Biswas, "Extraction of network information - quality and quantity - from nodes of neuronal network." Accepted at 2016 fifteenth IEEE International Conference on Information Technology, 2016. [13] S. Biswas, "How diverse is the network information obtained from the nodes of a biological neural network ?" Accepted at 2017 Eleventh IEEE International Conference on Intelligent Systems and Control, 2017. [14] S. Biswas, "Novel algorithm to spatially locate information and activity hub in biological neuronal network." Accepted at 2017 Fourth IEEE International Conference on Signal Processing and Integrated Networks, 2017. [15] E. Bullmore and O. Sporns, "Complex brain networks: graph theoretical analysis of structural and functional systems," Nature Reviews Neuro- science, vol. 10, no. 3, pp. 186–198, 2009. [16] D. A. Forsyth and J. Ponce, Computer vision: a modern approach. Prentice Hall Professional Technical Reference, 2002. [17] S. E. Umbaugh, Computer Vision and Image Processing: A Practical Approach Using Cviptools with Cdrom. Prentice Hall PTR, 1997. [18] C. M. Bishop, "Pattern recognition," Machine Learning, vol. 128, 2006. [19] B. Thakur, A. Mukherjee, A. Sen, and A. Banerjee, "A dynamical framework to relate perceptual variability with multisensory information processing," Scientific Reports, vol. 6, 2016. [20] M. E. J. Obien, K. Deligkaris, T. Bullmann, D. J. Bakkum, and U. Frey, "Revealing neuronal function through microelectrode array recordings," Frontiers in neuroscience, vol. 8, p. 423, 2015. [21] T. Watanabe and M. C. Raff, "Rod photoreceptor development in vitro: intrinsic properties of proliferating neuroepithelial cells change as development proceeds in the rat retina," Neuron, vol. 4, no. 3, pp. 461–467, 1990. [22] J. B. George, G. M. Abraham, K. Singh, S. M. Ankolekar, B. Amrutur, and S. K. Sikdar, "Input coding for neuro-electronic hybrid systems," Biosystems, vol. 126, pp. 1–11, 2014.
1210.7165
1
1210
2012-10-26T15:15:08
Cellular Adaptation Accounts for the Sparse and Reliable Sensory Stimulus Representation
[ "q-bio.NC", "physics.bio-ph" ]
Most neurons in peripheral sensory pathways initially respond vigorously when a preferred stimulus is presented, but adapt as stimulation continues. It is unclear how this phenomenon affects stimulus representation in the later stages of cortical sensory processing. Here, we show that a temporally sparse and reliable stimulus representation develops naturally in a network with adapting neurons. We find that cellular adaptation plays a critical role in the transient reduction of the trial-by-trial variability of cortical spiking, providing an explanation for a wide-spread and hitherto unexplained phenomenon by a simple mechanism. In insect olfaction, cellular adaptation is sufficient to explain the emergence of the temporally sparse and reliable stimulus representation in the mushroom body, independent of inhibitory mechanisms. Our results reveal a computational principle that relates neuronal firing rate adaptation to temporal sparse coding and variability suppression in nervous systems with a sequential processing architecture.
q-bio.NC
q-bio
1 Cellular Adaptation Accounts for the Sparse and Reliable Sensory Stimulus Representation Farzad Farkhooi1,∗, Anja Froese2, Eilif Muller3, Randolf Menzel2, Martin P. Nawrot1 1 Neuroinformatics & Theoretical Neuroscience, Freie Universitat Berlin, and Bernstein Center for Computational Neuroscience Berlin, Berlin, Germany. 2 Institute fur Biologie-Neurobiologie, Freie Universitat Berlin, Berlin, Germany. 3 Blue Brain Project, ´Ecole Polytechnique F´ed´erale de Lausanne, Lausanne, Switzerland. ∗ Corresponding author: Konigin-Luise-Strasse 1-3, 14195 Berlin, Germany [email protected] Abstract Most neurons in peripheral sensory pathways initially respond vigorously when a preferred stimulus is presented, but adapt as stimulation continues. It is unclear how this phenomenon affects stimulus representation in the later stages of cortical sensory processing. Here, we show that a temporally sparse and reliable stimulus representation develops naturally in a network with adapting neurons. We find that cellular adaptation plays a critical role in the transient reduction of the trial-by-trial variability of cortical spiking, providing an explanation for a wide-spread and hitherto unexplained phenomenon by a simple mechanism. In insect olfaction, cellular adaptation is sufficient to explain the emergence of the temporally sparse and reliable stimulus representation in the mushroom body, independent of inhibitory mechanisms. Our results reveal a computational principle that relates neuronal firing rate adaptation to temporal sparse coding and variability suppression in nervous systems with a sequential processing architecture. Introduction The phenomenon of spike-frequency adaptation (SFA) [1] (also known as spike-rate adaptation) is a fundamental process in nervous systems, which attenuates the neuronal stimulus responses to a lower level following an initial high firing. This process can be mediated by different cell-intrinsic mechanisms that involve a spike-triggered self-inhibition, and which can operate in a wide range of time scales [3,25]. These processes are probably related to the early evolution of the excitable membrane [36,38] and are common to vertebrate and invertebrate neurons, both in the peripheral and central nervous system [49]. Nonetheless, the functional consequences of SFA in peripheral stages of sensory processing on the representation of stimuli in later network stages remain unclear. For instance, while olfactory stimuli generally display a slow kinetics, the early olfactory system shows temporally adaptive and precise onset responses [19, 40], that may translate into narrow integration windows for the principal neurons in the next processing stage, the piriform cortex pyramidal cells in rodents [35] and the mushroom body Kenyon cells in insects [34]. However, it remains unclear how sparse single-neuron responses can reliably encode information in face of the response variability [41] and the sensitivity of cortical networks to small perturbations [24]. Here, we show that the SFA mechanism introduces a dynamical non-linearity in the transfer function of neurons. Subsequently, the response onset becomes progressively sparser when transmitted across successive processing stages. Additionally, the self-regulating effect of SFA causes a stimulus-triggered reduction of firing variability by modulating the average inhibition in the balanced cortical network. In this manner the temporally sparse representation is accompanied by an increased response reliability. Us- ing this theoretical framework on data obtained from the olfactory system of the honeybee, we show that 2 Figure 1. Neuronal adaptation in the multi-stage processing network. (A) Schematic illustration of a three-layered model of an adaptive pathway of sensory processing. The network consists of three consecutive adaptive populations. Each population receives sensory input from an afferent source (black arrows) and independent constant background excitation (dashed arrow). Input is modeled by a Gaussian density and a sensory stimulus presented to the first population is modeled by an increase in the mean input value. (B) Response profiles. The evoked state consists of a phasic-tonic response in all populations. The tonic response level is decremented across the consequtive populations. (C) Temporal sparseness Si average spike count at t = 500 ms in the first population. Responses become progressively sparser as the stimulus propagates into the network. (D) Secondary response profiles. The additional jump increase in stimulus strength at t = 1000 ms during the evoked state of the first stimulus results in a secondary phasic response in all populations with an amplitude overshoot in the 2nd and 3rd population. r is measured by the integral over the firing rate and normalized by the the sequential effect of the SFA shapes the Kenyon cells' temporal sparseness independent of inhibitory synaptic mechanisms. Our results reveal a generic, biophysically plausible mechanism that can explain the emergence of a reliable and sparse stimulus representation, indicating the importance of cellular adaptation in sensory computation at the network level. RESULTS Temporal sparseness emerges in successive adapting populations To examine how successive adapting populations can achieve temporal sparseness, first we mathematically analyzed a sequence of neuronal ensembles (Figure 1A), where each ensemble exhibits a generic model of mean firing rate adaptation by means of a slow negative self-feedback [3, 22, 29] (Materials and Pop. 1Pop. 2Pop. 3ABCD 3 Methods). This sequence of neuronal ensembles should be viewed as caricature for distinct stages in the pathway of sensory processing. For instance in the mammalian olfactory system the sensory pathway involves several stages from the olfactory sensory neurons to the olfactory bulb, the piriform cortex, and then to higher cortical areas (Figure 1A). The mean firing rate in the steady-state of a single adaptive population can be obtained by solving the rate consistency equation, ¯ri = f (µi − τsqs¯riσ2 i ), where ¯r is the equilibrium mean firing rate of the population i, µi and σ2 i are mean and variance of the total input into the population, respectively, f is the response function (input-output transfer function, or f I-curve) of the population mean activity, qs is the quantal conductance of the adaptation mechanism per unit of firing rate, and τs is the firing rate relaxation time constant [3, 22, 29]. It is known that any sufficiently slow modulation (1/rmax < τs) linearizes the steady-state solution, ¯ri, due to the self-inhibitory feedback being proportional to the firing rate [10]. Here, for simplicity, we studied the case where all populations in the network exhibit the same initial steady-state rate. This is achieved by adjustment of a constant background input to population i (doted arrows, Figure 1A), resembling the stimulus irrelevant interactions in the network. All populations are coupled by the same strength g. First, we calculated the average firing rate dynamics of the populations' responses following a step increase in the mean input to the first layer (black arrow, Figure 1A). By solving the dynamics of the mean firing rate and adaptation level concurrently, we obtained the mean- field approximation of the populations' firing rates (Materials and Methods). The responses of each population consisted of a phasic and a tonic part, typical for adapting neurons. Here, we plotted the mean firing rate of three consecutive populations (Figure 1B). The phasic response to a step increase in the input is presented across stages. However, the tonic response becomes increasingly suppressed in the later stages (Figure 1B). This phenomenon is a general feature of successive adaptive neuronal populations. To understand why the tonic response part, after the adaptation was reinstalled, is increasingly suppressed in later stages we used the linearity of the adaptive transfer function and its stability conditions at the steady-state [10]. Thus, a change in the mean rate of population i is the solution of ∆¯ri = f (g∆¯ri−1 − ∆¯riτsqs), and since the transfer function, f , is assumed to be increasing and the slope of ¯r is less than unity, the solution can only existe for ∆¯ri−1 > ∆¯ri (Materials and Methods). The result in this sub-section (Figure 1B-C) was established with a current based leaky integrated-and-fire response function. However, the analysis presented here extends to the majority of neuronal transfer functions since the stability and linearity of the adapted steady-states are granted [10, 22]. This simple effect leads to a progressively sparser representation across successive stage of a generic feed-forward adaptive processing. We quantified the temporal sparseness by Si 0 ri(t)dt, where ri(t) is the mean firing rate of population i and T is the observation time window. Normalization of this measure by the first population S1 r (T = 500) indicates sparser responses in later stages of the adaptive network (Figure 1C). Does the suppression of the adapted response level impair the information about the presence of the stimulus? To explore this, we studied a secondary increase in stimulus strengths of equal magnitude after 1 second when the network has relaxed to the evoked equilibrium (Figure 1D). The secondary stimulus jump induced a secondary phasic response of comparable magnitude in the first population (Figure 1D). However, in the later populations this jump evoked an icreased peak rate in the phasic response (Figure 1D). Notably, the coupling factor g between the populations shapes this phenomena. Here, we adjusted g to achieve an equal onset response magnitude across the populations for the first stimulus jump at t = 0, and a slight increase in the population onset response in the first population is amplified in the later stages. This is due to the fact that the later stages accumulated less adaptation in their evoked steady-state (level of adaption is proportional to mean firing rate). Importantly, this result confirms that the sustained presence of the stimulus is indeed stored in the level of cellular adaptation [32], even though it is not reflected in the firing rate of the last population. Therefore, regardless of the absolute amplitude of responses, the relative relation between secondary and initial onset keeps increasing across layers. This type of a secondary overshoot is also experimentally known as sensory sensitization, where r(T ) =(cid:82) T 4 an additional increase in the stimulus strength significantly enhances the responsiveness of later stages after the network converged to an adapted steady-state [18]. Adaptation increases response reliability in the cortical network The mean firing rate approach as above is insufficient to determine how reliable the observed response transients are across repeated stimulations. In the prevailing spiking network model of the cortex, the bal- ance of excitation and inhibition is quickly reinstalled within milliseconds after the onset of an excitatory input [47] and self-generating recurrent fluctuations strongly dominate the dynamics of the interactions. It has thus been questioned whether a sparse response of only very few action potentials could reliably encode the presence of a stimulus [24]. To investigate the reliability of adaptive mapping from a dense stimulus to a sparse cortical spkie response across successive processing stages we employed an adaptive population density formalisms [11, 29] (Materials and Methods) along with numerical network simulations. We embeded a two- layered sensory network with an afferent ensemble projecting to a cortical network (Figure 2A). The afferent ensemble consisted of 4,000 adaptive neurons that included voltage dynamics, conductance-based synapses, and spike-induced adaptation [29]. It resembles the sub-cortical sensory processing and each neuron in the afferent ensemble projects to randomly 1% of the neurons in the cortical network. This is a large circuit of the balanced network (Figure 2A) with 10,000 excitatory and 2,500 inhibitory neurons with a typical random diluted connectivity of 1%. The spiking neuron model in the cortical network again includes voltage dynamics, conductance-based synapses, and spike-induced adaptation [29]. All neurons are alike and parameters are given in Table.3 in Muller et al. [29]. With appropriate adjustment of the synaptic weights, the cortical network operates in a globally balanced manner, producing irregular, asynchronous activity [20, 47, 48]. The distribution of firing rates for the network approximates a power- law density [37] with an average firing rate of ≈ 3.0 Hz (Figure 2B) and the coefficients of variation (Cv) for the inter-spike intervals is centered at a value slightly greater than unity (Figure 2C) indicating the globally balanced and irregular state of the network [48]. Noteworthy, the activity of neurons in both stages is fairly incoherent and spiking in each sub-network is independent. Therefore, one can apply an adiabatic elimination of the fast variables and formulate a population density description where a detailed neuron model reduces to a stochastic point process [11,29] that provides an analytical approximation and understanding of the simulation results in this section (Materials and Methods). The background input is modeled as a set of independent Poisson processes that drive both sub- networks (dashed arrows, Figure 2A). The stimulus dependent input is an increase in the intensity of the Poisson input into the afferent ensemble (solid arrow, Figure 2A). Before the stimulus became active at time t = 0, a typical neuron showed an irregular spiking activity in both network stages (Figure 2B- D). However, when a sufficiently strong stimulus is applied, neurons in both stages exhibited a transient response before the population mean firing rate converges back to a new level of steady-state (Figure 2D,E). The population firing rate in the numerical simulations (solid line, Figure 2E) follow well the adaptive population density treatment (filled circles, Figure 2E). To measure the effect of neuronal adaptation on the temporal sparseness, we again computed the number of spikes per neuron after the stimulus onset, Se(T ). We compare our standard adaptive network with an adaptation time constant of τs = 110ms (solid lines, Figure 2F) with an only weakly-adaptive control network (τs =30ms; dashed lines, Figure 2F). Note, that the adaptation time constant in the weakly adaptive network is about equal to the membrane time constant and therefore plays a minor role for the network dynamics. It showed that both sub-networks generated sharp phasic response, which in the case of the cortical network evoked a single sharply timed spike within the first ≈ 10 − 20 ms in a subset of neurons (Figure 2B,F). In the control case, the response is non-sparse and response spikes are distrubuted throughout the stimulus period (dashed lines, Figure 2F). Overall, strong adaptation reduces the total number of stimulus-induced action potentials per neuron and concentrates their occurrence within an initial brief phase following fast changes in the stimulus. Thus, in accordance with the results 5 of the rate-based model in the previous section, one can conclude that the sequential stimulation of adaptive network stages can account for the emergence of a sparse stimulus representation in a cortical population. To reveal the effect of adaptation on the response variability, we employed the time-resolved Fano factor [31], F , which measures the spike-count variance divided by the mean spike count across 200 repeated stimulations. Spikes were counted in a 50ms time window and a sliding of 10ms [9]. As before, we compared our standard adaptive network (Figure 2G,I; τs = 110ms) with the control network (Figure 2H,J; τs=30ms). Since the Fano factor is known to be strongly dependent on the firing rate, we adjusted the input to the latter such that the steady state firing rates in both networks were mean-matched [9]. The input Poisson spike trains (F = 1) translated into slightly more regular spontaneous (t < 0) activity in the afferent ensemble (Figure 2G), as neuronal membrane filtering and refractoriness reduced the output variability. After the stimulus onset (t = 0), due to the increase in the mean input rate, the average firing rate increased, however the variance of the number of events per trial did not increase proportionally. Therefore, we observed a reduction in the Fano factor (Figure 2G). This phenomena is independent of the adaptation mechanism in the neuron model and a quantitatively similar reduction can be observed in the weakly adaptive afferent ensemble (Figure 2H). A comparison between our standard adaptive and the control case reveals that the adaptive network is generally more regular in the background and in the evoked state (Figure 2G,H). This is due to the previously known effect, where adaptation induces negative serial correlations in the inter-spike intervals [5, 11] and as a result reduces the Fano factor [7, 11]. In the next stage of processing, the distribution of F across neurons during spontaneous activity is high due to the self-generated noise of the balanced circuits [23, 47]. This closely follows a wide spread experimental finding where F > 1 in the spontaneous cortical activity [9] ( Figure 2G,H). Whenever a sufficiently strong stimulus was applied, the internally generated fluctuations in the adaptive balanced network were transiently suppressed, and as a result the Fano factor dropped sharply (Figure 2I). However, this reduction of the Fano factor is a temporary phenomenon and F converges back to slightly above the baseline variability (Figure 2I). At the same time, the evoked steady-state firing returned back to the irregular and asynchronous state (Figure 2D). Indeed, this effect corresponds to a temporally mis-match in the balanced input conditions to the cortical neurons since the self-inhibitory and slower adaption effect prevents a rapid adjustment to the new input regime. This can be observed in the time course of variability suppression that closely reflects the time constant of adaptation (Figure 2I). The afferent ensemble structured the input to the cortical ensemble, determining the magnitude of the observed reduction effect, and under the control condition where a pure Poisson input is provided to the cortical balanced network, the reduction in F is reduced but the time scale of recovery remains unaltered (crosses, Figure 2I). We contrast this adaptive behavior with the variability dynamics in the weakly adaptive balanced network (Figure 2J). In this case there is no reduction in F , because for a short adaptation time constant the convergence to the balanced state is very rapid [47]. The small increase in the input noise strength leads to an increase of the self-generated randomness of the balanced network [23, 28]. To further understand the effect of the recurrent connectivity on the variability, one can compare the evoked variability between the afferent ensemble and the cortical network. In the latter, fluctuation intrinsic to the cortical networks dynamically bring back the cortical circuits to the spontaneous level of high variability within the time scale of adaption relaxation (Figure 2I). This suggests the transient suppression of variability during a stimulus response as a cortical-wide effect [9] could be caused by slow adjustment of self-regulatory feedback. 6 Adaptive networks generate sparse and reliable responses in the insect olfac- tory system As a case study, we investigated the role of adaptation for the emergence of the sparse temporal code in the insects olfactory system, which is analogoue to the mammalian olfactory system. We simulated a reduced generic model of olfactory processing in insects using the adaptive neuron model [29]. The model network consisted of an input layer with 1,480 olfactory sensory neurons (OSNs), which project to the next layer representing the antennal lobe circuit with 24 projection neurons (PNs) and 96 inhibitory local inter-neurons (LNs) that form a local feed-forward inhibitory micro-circuit with the PNs. The third layer holds 1,000 Kenyon cells (KCs) receiving divergent-convergent input from PNs. The relative numbers for all neurons approximate the anatomical ratios found in the olfactory pathway of the honeybee [27] (Figure 3A). We introduced heterogeneity among neurons by randomizing their synaptic time constants and the connectivity probabilities are chosen according to anatomical studies. Synaptic weights were adjusted to achieve spontaneous firing statistics that match the observed physiological regimes. The SFA parameters were identical throughout the network with τs = 110 ms (see Materials and Methods for details). Using this model, we sought to understand how adaptation contributes to temporally sparse odor representations in the KC layer in a small sized network and under highly fluctuating input conditions. We simulated the input to each OSN by an independent Poisson process, which is thought to be reminiscent of the transduction process at the olfactory receptor level [30]. Stimulus activation was modeled by a step increase in the Poisson intensity with uniformly jittered onset across the OSN population (Materials and Methods). Following a transient onset response the OSNs adapted their firing to a new steady-state (Figure 3B,C). The pronounced effect of adaptation becomes apparent when the adaptive population response is compared to the OSN responses in the control network without any adaptation (τs=0; dashed line, Figure 3C). In the next layer, the PN population activity is reflected in a dominant phasic-tonic response profile (green line, Figure 3D), which closely matches the experimental observation [19]. This is due to the self-inhibitory effect of the SFA mechanism, and to the feedforward inhibition received from the LNs (magenta line, Figure 3D). Consequently, the KCs in the third layer produced only very few action potentials following the response onset (red line , Figure 3E,G). The average number of emitted KC response spikes per neuron, Se, is small in the adaptive network (average < 2) whereas KCs continue spiking throughout stimulus presentation in the non-adaptive network (Figure 3G). This finding closely resembles experimental findings of temporal sparseness of KC responses in different insect species [6,34,43] and quantitatively matches the KC response statistics provided by Ito and colleagues [15]. The simulation results obtained here confirm the mathematically derived results in the first results section (Figure 1) and show that neuronal adaption can cause a temporal sparse representation even in a fairly small and highly structured layered network where the mathematical assumptions of infinite network size and fundamentally incoherent activity are not full filled. To test the effect of inhibition in the LN-PN micro-circuitry within the antennal lobe layer on the emergence of temporal sparseness in the KC layer, we deactivated all LN-PN feedforward connections and kept all other parameters fix. We found a profound increase in the amplitude of the KC popu- lation response, both in the adaptive (red line,Figure 3F) and the non-adaptive network (dashed red line,Figure 3F). This increase in response amplitude is carried by an increase in the number of respond- ing KCs due to the increased excitatory input from the PNs, implying a strong reduction in the KC population sparseness. Importantly, removing local inhibition did not alter the temporal profile of the KC population response in the adaptive network (cf. red lines in Figure 3E,F), and thus temporal sparseness was independent of inhibition in our network model. How reliable is the sparse spike response across trials in single KCs? To answer this question, we again measured the robustness of the stimulus representation by estimating the Fano factor across 200 simulation trials (Figure 3H). The network with adaptive neurons and inhibitory micro-circuitry ex- hibited a low Fano factor (median ≈ 0.3) and a narrow distribution across all neurons. This follows the 7 experimental finding that the few spikes emitted by KCs are highly reliably [15] (network 1, Figure 3H). Turning off the inhibitory micro-circuitry did not significantly change the response reliability (Wilcoxon rank sum test, p-value = 0.01; network 2, Figure 3H). However, both networks that lacked adaptation exhibited a significantly increased variability with a median Fano factor close to one (Wilcoxon rank sum test, p-value = 0.01; networks 3 and 4, Figure 3H), independent of the presence or absence of inhibition within the antennal lobe. To explore whether neuronal adaptation could be the responsible mechanism for temporal sparseness in the biological network, we performed a set of Calcium imaging experiments, monitoring Calcium responses in the KC population of the honeybee mushroom body [43] (Materials and Methods). Our computational model (Figure 3) predicted that blocking of the inhibitory microcircuit would increase the population response amplitude but should not alter the temporal dynamics of the KC population response. In a set of experiments, we tested this hypothesis by comparison of the KCs' evoked activity in the presence and absence of GABAergic inhibition (Materials and Methods). First, we analyzed the normalized Calcium response signal within the mushroom body lip region in response to a 3s, 2s, 1s and 0.5s odor stimulus (Figure 4A). We observed the same brief phasic response following stimulus onset in all four cases with a characteristic slope of Calcium response decay that has been reported previously [43]. Bath application of the GABAA antagonist picrotoxin (PTX) did not change the time course of the Calcium response dynamics (Figure 4B-D). The effectiveness of the drug was verified by the increased population response amplitude in initial phase (Figure 4C). Next, we tested the GABAB antagonist hydrochloride (CGP) using the same protocol and again found an increase in the response magnitude but no alteration of the response dynamics (Figure 4E,F). DISCUSSION We here propose that a simple neuron-intrinsic mechanism of spike-triggered adaptation can account for a reliable and temporally sparse sensory stimulus representation across stages of sensory processing. At the single neuron level SFA is known to induce the functional property of a fractional differentiation with respect to the temporal profile of the input and thus offers the possibility of tuning the neuron's response properties to the relevant stimulus time-scales at the cellular level [3, 4, 25]. Our results indicate that a sensory processing in a feedforeward network with adaptive neurons focuses on the temporal changes of the sensory input in a precise and temporally sparse manner (Figure 1B; Figure 2E and Figure 3E). At the same time the constancy of the stimulus is memorized in the cellular level of adaptation [32] (Figure 1D). We hypothesize that at later processing stages in the sensory pathway, notably in the sensory cortices and in the insect mushroom body, it is most relevant to encode and process relevant dynamic changes in the sensory environment and to neglect the static stimulus features as well as the dynamic fluctuations of receptor sensations present at very fast time scales. One prominent example is primate vision where, in the absence of the self-generated dynamics of retina input due to microsaccades, observers become functionally blind to stationary objects during fixations [26]. The emergence of a sparse representation has been demonstrated in various cortical sensory areas, for example in visual [44], auditory [14], somatosensory [16], and olfactory [35] cortices, and thus manifests a principle of sensory computation across sensory modalities and independent of the natural stimulus kinetic. However, it has been repeatedly questioned whether a few informative spikes can survive in the cortical network, which is highly sensitive to small perturbations [24, 50]. Our results show that a biologically realistic cellular mechanism implemented at successive network stages can transform a dense and highly variable Poisson input at the periphery into a sparse and highly reliable ensemble representation in the cortical network. These results reflect previous theoretical evidence that SFA has an extensive synchronizing-desynchronizing effect on population responses in a balanced network [46]. However, cellular adaptation has largely been neglected in theory and simulation of cortical networks, with very few exceptions [46], although it facilitates a transition from a dense rate code to a temporal 8 code expressed in the concerted spiking of cortical cell assemblies [21] (Figure 2D). Importantly, the adaption level adjusts with a dynamics that is slow compared to the dynamics of excitatoary and inhibitory synaptic inputs. This circumstance allows for a transient mismatch of the balanced state in the cortical network and thus leads to a transient reduction of the self-generated (recurrent) noise (Figure 2I). This, in turn, explains why the temporally sparse representation can be highly reliable. This result seems counter-intuitive in the light of previous studies that consistently stressed the persistence of noise in the balanced network [23, 24, 28]. However, cellular adaptation has largely been neglected in theory and simulation of cortical networks with very few exceptions [46]. Our results provide an understanding of the role of slow cellular adaptation dynamics for the transient suppression of the trial-by-trial response variability. We thus arrived at a coherent model explanation for the previously unexplained and wide-spread experimental phenomenon of a stimulus-driven transient reduction of response variability in cortical neurons, which has been observed in different species and across many cortical areas, from occipital to frontal, as e.g. reported by Churchland and colleagues for 14 independent datasets [9]. The insect olfactory system is experimentally well investigated and examplifies a pronounced sparse coding scheme at the level of the mushroom body KCs. The olfactory system is analog in invertebrates and vertebrates and the sparse stimulus representation is likewise observed in the pyramidal cells of the piriform cortex [35], and the rapid responses in the mitral cells in the olfactory bulb [40] compare to those of projection neurons in antennal lobe [19]. Our adaptive network model, designed in coarse analogy to the insect olfactory system, produced increasingly phasic population responses as the stimulus-driven acitivty propagated through the network. Our model results closely match the repeated experimental observation of temporally sparse KC responses in extracellular recordings from the locust [34] and manduca [6, 15], and in Calcium imaging in the honeybee [43]. In our experiments we could show that systemic blocking of GABAergic transmission did not affect the temporal sparseness of the KC population response in the honeybee (Figure 4). This result might seem to contradict former studies that stressed the role of inhibitory feed-forward [2] or feedback inhibition [13,33] for the emergence of KC sparseness. However, the suggested inhibitory mechanisms and the sequential effect in the adaptive network proposed here are not mutually exclusive and may act in concert to establish and maintain a temporal and spatial sparse code in a rich and dynamic natural olfactory scene. The adaptive network model manifests a low trial-to-trial variability of the sparse KC responses that typically consist of only 1-2 spikes. In consequence, a sparsly activated KC ensemble is able to robustly encode stimulus information. The low variability at the single cell level (Figure 3H) carries over to a low variability of the population response [11]. This benefits downstream processing in the mushroom body output neurons that integrate converging input from many KCs [27], and which were shown to reliably encode odor-reward associations in the honeybee [42]. Our results here bear consequences of general importance for sensory coding theories. A mechanism of self-inhibition at the cellular level can facilitate a temporally sparse ensemble code but does not require a well adjusted interplay between the excitatory and inhibitory circuitry at the network level. This network effect is robust due to the distributed nature of the underlying mechanism, which acts independently in each single neuron. The regularizing effect of self-inhibition increases the signal-to-noise ratio not only of single neuron responses but also of the neuronal population activity [11] that is postsynaptically integrated in downstream neurons. Materials and Methods Rate model of generic feedforward adaptive network To address analytically the sequential effect of adaptation in a feedforward network, we consider a model in which populations are described by their firing rates. Although firing rate models typically provide a fairly accurate description of network behavior when the neurons are firing asynchronously [45], they do not capture all features of realistic networks. Therefore, we verify all of our predictions with a population density formalism [11] as well as a large-scale simulation of realistic spiking neurons. To determine the mean activity dynamics of a consecutive populations, we employed an standard mean firing rate model of population i as (cid:40) ri = −ri + f (gri−1 + µi − aiσi) ai = − 1 ai + riqs τs 9 (1) ¯ri = f (mi − τsqs¯riσi). where f is the transfer input-output function, τs is the adaptation time-scale, g is the coupling factor between two populations and a is the adaptive negative feedback for the population i with qs strength and σ is the standard deviation of the input. For simplicity we define mi = gri−1 + µi. Given f the equilibrium can be determined by (2) This fix point is always stable [22] and the condition for the stability reads ∂rif (m − τsqs¯riσi)¯ri < 1 and ∂mi f (m − τsqs¯riσi)¯r < 0. It is also known that the ¯ri is a linear function, given a sufficiently slow adaptation τs and a non-linear shape of f [10]. Thus, if ri−1 = ri in two subsequent populations, and increased input to population i − 1 leads to smaller increase for next population i and the adapted level of responses satisfy ∆ri−1 > ∆ri. The stimilus onset response lies between the adapted steady-state and not adapted responses function f given the level of new input and can be analytical calculated accordingly [3]. Population density approach to the adaptive neuronal ensemble In Muller et al. [29] it is shown that by an adiabatic elimination of fast variables a detailed neuron model including voltage dynamics, conductance-based synapses, and spike-induced adaptation reduces to a stochastic point process. Thus, we define an orderly point process with a hazard function argument with state variable x as Pr(N [t, t + ∆t) > 0x, t) hx(x, t) := lim ∆t→0+ ∆t . (3) where N [t, t + ∆t) is the number of events in ∆t. We assume the dynamic of the adaptation variable is x = − 1 τs x + δ(t − tk)qs, (4) (cid:88) k (cid:21) (cid:20) x Pr(x, t) − hx(x, t) Pr(x, t). τs where tk is the time of kth spike in the ensemble. Thus, the state variable distribution at time t in the ensemble is governed by a master equation of the form ∂t Pr(x, t) = −∂x + hx(x − qs, t) Pr(x − qs, t) (5) We solve the Eq. 5 with the help of the transformations ts = η(x) := −τ log(x/q) and ψ(ts) = η(η−1(ts + q)) numerically [29]. It turns out that indeed hx is the input-out transfer function of neurons in the network where its instantaneous parameters are give by the input statistics [29]. Here, we used the mean- field formalism developed by Lechner et al. [23] to approximately determine the averaged input within a standard balanced network, as the parameters of the hazard function hx. Therefore, the functional form of its solution in a compact form is Pr(tst) = ktΩ(tst) (6) s is the initial condition state of the system and kt is a constant defined by (cid:82) Pr(tst)dts = 1. where t0 Hence, it can be shown that the firing rate and the consistency equation of the ensemble is rt = hts (tst)Pr(tst)dts. 10 (7) (8) (cid:90) ∞ −∞ (cid:90) (cid:90) Now, by applying the techniques in Farkhooi et al [11], we define a joint probability density as ρn(tn, tn xt0 x, t) where an nth event occurs at time tn > t and the state variable is tn and state of adaptation joint density recursively x. We can write n + 1th event time ρn+1(tn+1, tn+1 x t0 x, t) = ρ(tn+1 − tn, tn+1 x tn x, t)ρn(tn, tn xt0 x, t)dtndtn x (9) To simplify the integral equations, we use Bra-Ket notation and thereafter, we derive the Laplace trans- form of the joint density by where ρ1(t1, t1 x, t) = ρ(t1, t1 st0 t0 s, t) = ρn(s, tn tn x, t)(cid:105) x, t) Next, we define the operator Pn(st), ρn+1(s, tn+1 xt0 x, t)ρ(s, tn+1 xt0 x x ρn(st) = (cid:104)1 Pn(st) Pr(ts, t)(cid:105) = (cid:104)1 Pn(st) Pr(tx, t)(cid:105). Now, by employing P (1, st) = 1/(µ1s2)[ρ2(st) − 2ρ1(st) + 1] as in [11], we derive (10) (11) P (n, st) = 1/(µ1s2)[ρn+1(st) − 2ρn(st) + ρn−1(st)], (12) where P (n, st) is the Laplace transform of the probability density of observing n events in a given time window. Now we derive k ρk(st) = (cid:104)1 P(st)/(I − P(st)) Pr(st)(cid:105) (13) Ast =(cid:80) (cid:90) T where I is the identity operator. This equation represents the Laplace transfom of the autocorrelation function. Thus, the Fano factor is Jst = 1/µ1s2 − (1 + 2 Ast) and the inverse Laplace transform is JTt = 1 + (2/T ) (T − u)A(ut)du − T rt, (14) where A(ut) = L−1[ Ast]. 0 Computational model of insect olfactory coding Our model neuron is a general conductance-based integrate-and-fire neuron with spike-frequency adap- tation as it is proposed in Muller et al. [29]. The model phenomenologically captures a wide array of biophysical spike-frequency mechanisms such as M-type current, afterhyperpolarization (AHP-current) and even slow recovery from inactivation of the fast sodium current [29]. The model neuron is also known to perform high-pass filtering of the input frequencies following the universal model of adaptation [3]. Neuron parameters used follow the Table.3 in Muller et al. [29] The conductance model used for the static synapses between the neurons is alpha-shaped with gamma distributed time constants from Γ(2.5, 4) and Γ(5, 4) for excitatory and inhibitory synapses, respectively. All simulations were performed using the NEST simulator [12] version 2.0beta and the Pynest interface. 11 The network connectivity is straight forward: each PN and LN receives excitatory connections from 20% randomly chosen OSNs [8, 39]. Additionally, every PN receives input from 50% randomly chosen inhibitory LNs [8, 39]. In our model the PNs do not excite one another and each PN output diverges to 50% randomly chosen KCs [2, 17, 43]. We tuned the simulated network by adjusting the synaptic weights to achieve the same spontaneous firing rate as reported experimentally: OSNs 15-25Hz [30], LNs 4-10Hz [8], PNs 3-10 [19] and KCs 0.3-1.0Hz [15]. Experimental methods Experiments were performed following the methods published in Szyszka et al [43]. In summary, foraging honeybees (Apis mellifera) were caught at the entrance of the hive, immobilized by chilling on ice, and fixed in a plexiglas chamber before the head capsule was opened for dye injection. We retrogradely stained clawed Kenyon cells (KC) of the median calyx, using the calcium sensor FURA-2 dextran (Molecular Probes, Eugene, USA) with a dye loaded glass electrode, which was pricked into KC axons projecting to the ventral median part of the α-lobe [43]. After dye injection the head capsule was closed, bees feed and kept in a dark humid chamber for several hours. The processing of imaging data was performed with custom written routines in IDL (RSI, Boulder, CO, USA). In summary, changes in the calcium concentration were measured as absolute changes of fluorescence: a ratio was calculated from the light intensities measured at 340nm and 380nm illumination and the background fluorescence before odor onset was subtracted leading to ∆F with F = F340/F380. Odor stimulation was preformed under a 20x objective of the microscope, the naturally occurring plant odor octanol (Sigma Aldrich, Germany), diluted 1:100 in paraffine oil (FLUKA, Buchs, Switzerland), was delivered to both antennae of the bee using a computer controlled, custom made olfactometer. To this, odor loaded air was injected into a permanent airstream resulting in a further 1:10 dilution. Stimulus duration was 3 seconds if not mentioned otherwise. The air was permanently exhausted. For GABA blockage, a solution of 150 µl GABA receptor antagonist dissolved in ringer for final concentration (10−5M picrotoxin (PTX, Sigma Aldrich, Germany) or 5x10−4M CGP54626 (CGP, Tocris Bioscience, USA)) was bath applied to the brain after pre-treatment measurements. Measurements started 10 min after drug application. Acknowledgments We wish to thank Peter Latham and Yifat Prut for helpful comments on this manuscript. Generous funding was provided by the Bundesministerium fur Bildung und Forschung (Grant No.01GQ0941) to the Bernstein Focus Neuronal Basis of Learning (BFNL) and by the Deutsche Forschungsgemeinschaft (DFG) to the Collaborative Research Center for Theoretical Biology (SFB 618) and the DFG grant to R.M. and A.F. (Me 365/31-1). References 1. E. D. Adrian. The impulses produced by sensory nerve endings. The Journal of physiology, 61(1):4972, 1926. 2. Collins Assisi, Mark Stopfer, Gilles Laurent, and Maxim Bazhenov. Adaptive regulation of sparse- ness by feedforward inhibition. Nat Neurosci, 10(9):1176 -- 84, September 2007. 3. Jan Benda and Andreas V. M. Herz. A universal model for spike-frequency adaptation. Neural Computation, 15(11):2523 -- 2564, November 2003. 12 4. Jan Benda, Andre Longtin, and Len Maler. Spike-frequency adaptation separates transient com- munication signals from background oscillations. J. Neurosci., 25(9):2312 -- 2321, March 2005. 5. Jan Benda, Leonard Maler, and Andr Longtin. Linear versus nonlinear signal transmission in neuron models with adaptation currents or dynamic thresholds. Journal of Neurophysiology, 104(5):2806 -- 2820, November 2010. 6. Bede M Broome, Vivek Jayaraman, and Gilles Laurent. Encoding and decoding of overlapping odor sequences. Neuron, 51(4):467 -- 82, August 2006. 7. Maurice J. Chacron, Leonard Maler, and Joseph Bastian. Electroreceptor neuron dynamics shape information transmission. Nature Neuroscience, 8(5):673 -- 678, 2005. 8. Ya-Hui Chou, Maria L Spletter, Emre Yaksi, Jonathan C S Leong, Rachel I Wilson, and Liqun Luo. Diversity and wiring variability of olfactory local interneurons in the drosophila antennal lobe. Nature Neuroscience, 13(4):439 -- 449, February 2010. 9. Mark M Churchland, Byron M Yu, John P Cunningham, Leo P Sugrue, Marlene R Cohen, Greg S Corrado, William T Newsome, Andrew M Clark, Paymon Hosseini, Benjamin B Scott, David C Bradley, Matthew A Smith, Adam Kohn, J Anthony Movshon, Katherine M Armstrong, Tirin Moore, Steve W Chang, Lawrence H Snyder, Stephen G Lisberger, Nicholas J Priebe, Ian M Finn, David Ferster, Stephen I Ryu, Gopal Santhanam, Maneesh Sahani, and Krishna V Shenoy. Stimulus onset quenches neural variability: a widespread cortical phenomenon. Nat Neurosci, 13(3):369 -- 378, March 2010. 10. B. Ermentrout. Linearization of FI curves by adaptation. Neural computation, 10(7):17211729, 1998. 11. Farzad Farkhooi, Eilif Muller, and Martin P. Nawrot. Adaptation reduces variability of the neuronal population code. Physical Review E, 83(5):050905, May 2011. 12. Marc-Oliver Gewaltig and Markus Diesmann. NEST (NEural simulation tool). Scholarpedia, 2(4):1430, 2007. 13. N. Gupta and M. Stopfer. Functional analysis of a higher olfactory center, the lateral horn. Journal of Neuroscience, 32(24):8138 -- 8148, June 2012. 14. Tom Hromdka, Michael R DeWeese, and Anthony M Zador. Sparse representation of sounds in the unanesthetized auditory cortex. PLoS Biol, 6(1):e16, January 2008. 15. Iori Ito, Rose Chik-ying Ong, Baranidharan Raman, and Mark Stopfer. Sparse odor representation and olfactory learning. Nature Neuroscience, 11(10):1177 -- 1184, September 2008. 16. Shantanu P Jadhav, Jason Wolfe, and Daniel E Feldman. Sparse temporal coding of elementary tactile features during active whisker sensation. Nature Neuroscience, 12(6):792 -- 800, May 2009. 17. Ron A. Jortner, S. Sarah Farivar, and Gilles Laurent. A simple connectivity scheme for sparse coding in an olfactory system. The Journal of Neuroscience, 27(7):1659 -- 1669, February 2007. 18. Mikiko Kadohisa and Donald A. Wilson. Olfactory cortical adaptation facilitates detection of odors against background. Journal of Neurophysiology, 95(3):1888 -- 1896, March 2006. 19. Sabine Krofczik, Randolf Menzel, and Martin P. Nawrot. Rapid odor processing in the honey- bee antennal lobe network. Frontiers in Computational Neuroscience, 2, 2008. PMID: 19221584 PMCID: 2636688. 13 20. A. Kumar, S. Schrader, A. Aertsen, and S. Rotter. The high-conductance state of cortical networks. Neural Computation, 20(1):143, 2008. 21. Arvind Kumar, Stefan Rotter, and Ad Aertsen. Spiking activity propagation in neuronal networks: reconciling different perspectives on neural coding. Nat Rev Neurosci, 11(9):615 -- 627, 2010. 22. Giancarlo LaCamera, Alexander Rauch, Hans-R Luscher, Walter Senn, and Stefano Fusi. Min- imal models of adapted neuronal response to in vivo-like input currents. Neural Comput, 16(10):21012124, October 2004. 23. Alexander Lerchner, Cristina Ursta, John Hertz, Mandana Ahmadi, Pauline Ruffiot, and Sren Enemark. Response variability in balanced cortical networks. Neural Computation, 18(3):634 -- 659, 2006. 24. Michael London, Arnd Roth, Lisa Beeren, Michael Hausser, and Peter E. Latham. Sensitivity to perturbations in vivo implies high noise and suggests rate coding in cortex. Nature, 466(7302):123 -- 127, July 2010. 25. Brian N. Lundstrom, Matthew H. Higgs, William J. Spain, and Adrienne L. Fairhall. Fractional differentiation by neocortical pyramidal neurons. Nature Neuroscience, 11(11):1335 -- 1342, 2008. 26. Susana Martinez-Conde, Stephen L. Macknik, Xoana G. Troncoso, and David H. Hubel. Microsac- cades: a neurophysiological analysis. Trends in Neurosciences, 32(9):463 -- 475, September 2009. 27. R. Menzel and Larry R. Squire. Olfaction in invertebrates: Honeybee. In Encyclopedia of Neuro- science, pages 43 -- 48. Academic Press, Oxford, 2009. 28. Michael Monteforte and Fred Wolf. Dynamical entropy production in spiking neuron networks in the balanced state. Physical Review Letters, 105(26):268104, December 2010. 29. Eilif Muller, Lars Buesing, Johannes Schemmel, and Karlheinz Meier. Spike-frequency adapting neural ensembles: Beyond mean adaptation and renewal theories. Neural Comp., 19(11):2958 -- 3010, 2007. 30. Katherine I Nagel and Rachel I Wilson. Biophysical mechanisms underlying olfactory receptor neuron dynamics. Nature Neuroscience, 14(2):208 -- 216, January 2011. 31. Martin P. Nawrot, Clemens Boucsein, Victor Rodriguez Molina, Alexa Riehle, Ad Aertsen, and Stefan Rotter. Measurement of variability dynamics in cortical spike trains. Journal of Neuroscience Methods, 169(2):374 -- 390, April 2008. 32. William H. Nesse, Leonard Maler, and Andr Longtin. Biophysical information representation in temporally correlated spike trains. Proceedings of the National Academy of Sciences, 107(51):21973 -- 21978, December 2010. 33. Maria Papadopoulou, Stijn Cassenaer, Thomas Nowotny, and Gilles Laurent. Normalization for sparse encoding of odors by a wide-field interneuron. Science, 332(6030):721 -- 725, May 2011. 34. Javier Perez-Orive, Ofer Mazor, Glenn C Turner, Stijn Cassenaer, Rachel I Wilson, and Gilles Laurent. Oscillations and sparsening of odor representations in the mushroom body. Science, 297(5580):359 -- 65, July 2002. 35. Cindy Poo and Jeffry S. Isaacson. Odor representations in olfactory cortex: Sparse coding, global inhibition, and oscillations. Neuron, 62(6):850 -- 861, June 2009. 14 36. R. Ranganathan. Evolutionary origins of ion channels. Proceedings of the National Academy of Sciences of the United States of America, 91(9):3484, 1994. 37. Alex Roxin, Nicolas Brunel, David Hansel, Gianluigi Mongillo, and Carl van Vreeswijk. On the dis- tribution of firing rates in networks of cortical neurons. The Journal of Neuroscience, 31(45):16217 -- 16226, November 2011. 38. B. Rudy. Diversity and ubiquity of k channels. Neuroscience, 25(3):729 -- 749, June 1988. 39. Silke Sachse and C. Giovanni Galizia. Role of inhibition for temporal and spatial odor representation in olfactory output neurons: A calcium imaging study. J Neurophysiol, 87(2):1106 -- 1117, February 2002. 40. Roman Shusterman, Matthew C Smear, Alexei A Koulakov, and Dmitry Rinberg. Precise olfactory responses tile the sniff cycle. Nature Neuroscience, 14(8):1039 -- 1044, July 2011. 41. Richard B. Stein, E. Roderich Gossen, and Kelvin E. Jones. Neuronal variability: noise or part of the signal? Nat Rev Neurosci, 6(5):389 -- 397, May 2005. 42. Martin Fritz Strube-Bloss, Martin Paul Nawrot, and Randolf Menzel. Mushroom body output neurons encode odor-reward associations. J. Neurosci., 31(8):3129 -- 3140, February 2011. 43. Paul Szyszka, Mathias Ditzen, Alexander Galkin, C. Giovanni Galizia, and Randolf Menzel. Spars- ening and temporal sharpening of olfactory representations in the honeybee mushroom bodies. Journal of Neurophysiology, 94(5):3303 -- 3313, November 2005. 44. David J. Tolhurst, Darragh Smyth, and Ian D. Thompson. The sparseness of neuronal responses in ferret primary visual cortex. The Journal of Neuroscience, 29(8):2355 -- 2370, February 2009. 45. A. Treves. Mean-field analysis of neuronal spike dynamics. Network: Computation in Neural Systems, 4(3):259284, 1993. 46. C. van Vreeswijk. Analysis of the asynchronous state in networks of strongly coupled oscillators. Phys Rev Lett, 84(22):51105113, May 2000. 47. C. van Vreeswijk and H. Sompolinsky. Chaotic balanced state in a model of cortical circuits. Neural Comput, 10(6):1321 -- 71, August 1998. 48. Tim P Vogels and L F Abbott. Gating multiple signals through detailed balance of excitation and inhibition in spiking networks. Nat Neurosci, 12(4):483 -- 491, April 2009. 49. Barry Wark, Brian Nils Lundstrom, and Adrienne Fairhall. Sensory adaptation. Current opinion in neurobiology, 17(4):423 -- 429, August 2007. PMID: 17714934 PMCID: PMC2084204. 50. Jason Wolfe, Arthur R Houweling, and Michael Brecht. Sparse and powerful cortical spikes. Current Opinion in Neurobiology, 20(3):306 -- 312, June 2010. 15 Figure 2. Reliability of a temporally sparse code in the balanced cortical network. (A) Schematic of a two-layer model of sub-cortical and early cortical sensory processing. The afferent ensemble (blue) consists of 4,000 independent neurons, and each neuron projects to 1% of the neurons in the cortical network (green). The cortical network is a balanced network in the asynchronous and irregular state with random connectivity. In both populations black circles represent excitatory neurons and magenta circles represent inhibitory cells. (B) The distribution of firing rates across neurons in the cortical network is fat-tailed and the average firing rate is approximately 3Hz. (C) The distribution of the coefficient of variation (Cv) across neurons in the balanced cortical network confirms irregular spiking. (D) Spike raster plot for a sample set of 30 afferent neurons (blue dots) and 30 excitatory cortical neurons (green dots). At t = 0 (gray triangle) the stimulus presentation starts. (E) Population averaged firing for both network stages. The simulation (solid lines) follows the calculated ensemble average predicted by the adaptive density treatment (black circles). The firing rate in the simulated network is estimated with a 20ms bin size. (F) Number of spikes per neuron Se after stimulus onset (t = 0) for the adaptive network (solid lines) and the weakly adaptive control network (dashed lines). Cortical excitatory neurons (green) produced less spikes than neurons in the earlier stage (afferent ensemble, blue). The shaded area indicates the standard deviation across neurons. (G,H) Fano factor dynamics of the afferent ensemble in the network with strongly adapting neurons (G, τs = 110 ms) and in the weakly-adaptive (h, τs = 30 ms) network, estimated across 200 trials in a 50ms window and a sliding of 10ms for the ensemble network with adaptation. The black circles indicate the theoretical value of the Fano factor computed by adaptive density treatment and shaded area is the standard deviation of the Fano factor across neurons in the network. (I) The Fano factor of strongly adaptive neurons in cortical balanced network reduced transiently during the initial phasic response part. The crosses show the adaptive cortical ensemble Fano factor for the case where the afferent ensemble neurons were modeled as a Poisson process with the same steady-state firing rate and without adaption. (J) The Fano factor in the weakly adaptive cortical network did not exhibit a reduction during stimulation. Poisson ProcessesAfferent ensembleCortical networkABCDEFGHIJ 16 Figure 3. Neuronal adaptation generates temporal sparseness in a generic model of the insect olfactory network. (A) Schematic drawing of a simplified model of the insect olfactory network for a single pathway of odor coding. Olfactory receptor neurons (OSNs, first layer, n=1,480) project to the antennal lobe network (second layer) consisting of projection neurons (PNs, n=24) and local neurons (magenta, n=96), which make inhibitory connections with PNs. PNs project to the Kenyon cells (KCs) in the mushroom body (third layer). (B) Spike raster plot of randomly selected OSNs (blue), LNs (magenta), PNs (green) and KCs (red) indicates that spiking activity in the network became progressively sparser as the Poisson input propagated into the network. (C) Average population rate of OSNs in the adaptive network (blue solid line) and the non-adaptive control network (dashed blue lines). The shaded area indicates the firing rate distribution of the neurons. The firing rate was estimated with 20ms bin size. (D) Average response in the antennal lobe network. PNs (green) and LNs (magenta) exhibited the typical phasic-tonic response profile in the adaptive network (solid lines) but not in the non-adaptive case (dashed lines). (E) Kenyon cell activity. In the adaptive network the KC population exhibits a brief response immideately after stimulus onset, which quickly returns close to baseline. This is contrasted by a tonic response profile throughout the stimulus in the non-adaptive case. (F) Effect of the inhibitory micro-circuit. By turning off the inhibitory LN-PN connections the population response amplitude of the KCs was increased, while the population response dynamics did not change.(G) Sparseness of KCs. The average number of spikes per neuron emitted since stimulus onset indicates that the adaptive ensemble encodes stimulus informatoin with only very few spikes. (G) Reliability of KCs responses. The Fano factor of the KCs in different network scenarios is estimated across 200 trials in a 100ms time window after stimulus onset. Network 1: (+)Adaptation (+)Inhibition, network 2: (+)Adaptation (-)Inhibition, network 3: (-)Adaptation (-)Inhibition, network 4: (-)Adaptation (+)Inhibition. Both networks with SFA are significantly more reliable in their stimulus encoding than the non-adaptive networks. Poisson ProcessesReceptor Neurons Antennal lobeKenyon CellsABCDEFGH 17 Figure 4. Blocking GABAergic transmission in the honeybee changes amplitude but not duration of the KC population response. (A) Temporal response profile of the Ca signal imaged in the mushroom body lip region of one honeybee for different stimulus durations as indicated by color. (B) Temporal response profiles as in (A) in one honeybee after application of PTX. (C) Response profiles imaged from 6 control animals (gray) and their average (black) for a 3s stimulus as indicated by the stimulus bar. The responses measured in 6 animals in which GABAA transmission was blocked with PTX (red) shows a considerably higher population response amplitude. The shaded area indicated the standard deviation of responses across bees. (D) Average amplitude-normalized responses are highly similar in animals treated with PTX and control animals. (E) Blocking GABAB transmission with CGP in 6 animals (blue) again results in an increased response amplitude compared to 6 control animals (black). The shaded area indicates the standard deviation across individuals. (F) Average normalized response profiles are highly similar in the CGP-treated and control animals. ABCDEF
1704.08669
1
1704
2017-04-27T17:33:32
Evolution of moments and correlations in non-renewal escape-time processes
[ "q-bio.NC", "math.PR", "physics.comp-ph", "physics.data-an" ]
The theoretical description of non-renewal stochastic systems is a challenge. Analytical results are often not available or can only be obtained under strong conditions, limiting their applicability. Also, numerical results have mostly been obtained by ad-hoc Monte--Carlo simulations, which are usually computationally expensive when a high degree of accuracy is needed. To gain quantitative insight into these systems under general conditions, we here introduce a numerical iterated first-passage time approach based on solving the time-dependent Fokker-Planck equation (FPE) to describe the statistics of non-renewal stochastic systems. We illustrate the approach using spike-triggered neuronal adaptation in the leaky and perfect integrate-and-fire model, respectively. The transition to stationarity of first-passage time moments and their sequential correlations occur on a non-trivial timescale that depends on all system parameters. Surprisingly this is so for both single exponential and scale-free power-law adaptation. The method works beyond the small noise and timescale separation approximations. It shows excellent agreement with direct Monte Carlo simulations, which allows for the computation of transient and stationary distributions. We compare different methods to compute the evolution of the moments and serial correlation coefficients (SCC), and discuss the challenge of reliably computing the SCC which we find to be very sensitive to numerical inaccuracies for both the leaky and perfect integrate-and-fire models. In conclusion, our methods provide a general picture of non-renewal dynamics in a wide range of stochastic systems exhibiting short and long-range correlations.
q-bio.NC
q-bio
Evolution of moments and correlations in non-renewal escape-time processes Wilhelm Braun,1, ∗ Rüdiger Thul,2 and André Longtin1 1Department of Physics and Centre for Neural Dynamics, University of Ottawa, 598 King Edward, Ottawa K1N 6N5, Canada 2Centre for Mathematical Medicine and Biology, School of Mathematical Sciences, University of Nottingham, Nottingham, NG7 2RD, UK (Dated: September 28, 2018) 7 1 0 2 r p A 7 2 ] . C N o i b - q [ 1 v 9 6 6 8 0 . 4 0 7 1 : v i X r a The theoretical description of non-renewal stochastic systems is a challenge. Analytical results are often not available or can only be obtained under strong conditions, limiting their applicability. Also, numerical results have mostly been obtained by ad-hoc Monte -- Carlo simulations, which are usually computationally expensive when a high degree of accuracy is needed. To gain quantitative insight into these systems under general conditions, we here introduce a numerical iterated first-passage time approach based on solving the time-dependent Fokker -- Planck equation (FPE) to describe the statistics of non-renewal stochastic systems. We illustrate the approach using spike-triggered neuronal adaptation in the leaky and perfect integrate-and-fire model, respectively. The transition to stationarity of first-passage time moments and their sequential correlations occur on a non-trivial timescale that depends on all system parameters. Surprisingly this is so for both single exponential and scale-free power-law adaptation. The method works beyond the small noise and timescale separation approximations. It shows excellent agreement with direct Monte Carlo simulations, which allows for the computation of transient and stationary distributions. We compare different methods to compute the evolution of the moments and serial correlation coefficients (SCC), and discuss the challenge of reliably computing the SCC which we find to be very sensitive to numerical inaccuracies for both the leaky and perfect integrate-and-fire models. In conclusion, our methods provide a general picture of non-renewal dynamics in a wide range of stochastic systems exhibiting short and long-range correlations. I. INTRODUCTION A general property of diverse systems, ranging from su- perconducting quantum interference devices (SQUIDs) [1], to lasers [2] to excitable cells [3 -- 6] is that time in- tervals between specific events are not statistically inde- pendent. The theoretical description of such non-renewal stochastic processes [7] poses a significant challenge, as it implies that the present state of the system depends, in general, on the whole past evolution or parts of it, and not just on the previous state. Analytical approximations to tackle such memory effects have included the assumption of stationarity [8], small stochasticity [9] and time-scale separation [10 -- 12] between stochastic and deterministic parts of the dynamics. Even if these approximations allow for some insight into the parameter dependence of e.g. serial correlations and can be used to understand experimental data, as exem- plified in [9, 13 -- 16] in the context of excitable systems, it is desirable to understand the statistics of model systems without making simplifying assumptions. Regarding sta- tionarity, real systems rarely operate in a stationary state due to transients that arise from deterministic or ran- dom perturbations. A prominent example are cortical neurons. An average cortical neuron receives random in- puts from approximately 104 other neurons, whose ac- tivity is modulated by non-stationary sensory and other inputs, resulting in transient neuronal dynamics [17] that ∗ [email protected] only become stationary after a certain time. It is there- fore important to understand how statistical properties of inter-event times evolve and become invariant following a transient regime due to internal dynamics and external inputs. Keeping with illustrations from neural dynam- ics, it is well known that physiologically relevant pro- cesses underlying neural coding rarely only have one well- defined timescale [18, 19]. This has lead researchers in various theoretical fields to consider multiple time-scale dynamics [20]. An important example of a system with multiple time scales is neuronal adaptation, where a neu- ron's firing rate adjusts in response to a stimulus. Adap- tation with multiple timescales, or even no time scale as in the case of power-law adaptation [21 -- 23], is now known to be biophysically relevant, and even optimal for some tasks [24]. Recently, it was also shown that a neuron model with adaptive firing thresholds exhibiting multi- ple timescales is the optimal choice for the prediction of spike times in cortical neurons [25, 26]. Therefore, a theoretical description of adaptation without a single well-defined timescale is an important goal. In this paper, we show how to describe two-dimensional non-renewal dynamics by an iterated first-passage time (iFPT) approach. This approach allows us to determine stationary statistical properties of the system as well as providing a description of the transition to stationarity. We furthermore show how to compute serial correlations in the time series generated by the firing times of the system. While our approach is general and applicable to any system where first-passage times [27] play a role, we illustrate its versatility with two important examples, namely spike-triggered neuronal adaptation with a sin- gle exponential current and a power-law current without an intrinsic timescale, respectively. Using the underly- ing time-dependent FPE to describe the system, we only need to apply mathematically convenient standard ab- sorbing boundary conditions to obtain stationary distri- butions, e.g. that of the adaptation current upon firing. Moreover, the methods developed here can easily be ex- tended to models with correlation-generating determin- istic input currents as recently considered in [28]. II. MODEL We consider a stochastic differential equation (SDE) driven by an external signal s(t): dX(t) = µ(X(t))dt + φ(X(t))dW (t) − s(t)dt . (1) X is defined on the domain (−∞, xth]. If X reaches xth, the system is said to have generated and event, and X is instantaneously reset to 0. For all examples in this study, we chose the Ornstein -- Uhlenbeck process (OUP) given its prominence in the field of stochastic systems. For the OUP, we fix the correlation time τm = 1 γ , bias current I0 and noise intensity σ as follows: µ(X(t)) = γ(I0 − X(t)), φ(X(t)) = σγ. W (t) is a standard Brownian motion and we set xth = 1. Given that the OUP is the basis for integrate-and-fire (IF) neuron models, which are among the most popular neuron descriptions [29], we refer to events as spikes and to s(t) as a time-dependent adapta- tion current in the present study. The general dynamics of s(t) obeys a single autonomous ordinary differential equation (ODE) s = ω(s) , (2) and s is increased by a fixed amount κ when X = xth: s → s + κ, which is the mechanism for spike-triggered adaptation [30]. When s(t) is also reset to its starting value s(0), the model is a renewal model and its firing statistics may be studied using standard techniques, see e.g. [31] for a recent review. Here we focus on two forms of the adaptation current. The first one is power-law adaptation, for which ω(s) = − 1 α s2(t) . (cid:16) t (3) (cid:17)−1 . This ODE has the general solution s(t) = Therefore, the current s in this case has a power-law time dependence with no intrinsic time scale [21]. The second adaptation current is given by a single expo- nential decay with time scale τa: α + 1 s(0) which has the general solution s(t) = s(0)e The time to the first spike event is the following first- passage time (FPT): τa . − t 2 T1 = inf(t > 0 : X(t) > xthX(0) = 0, s(t = 0) = s(0)) . FIG. 1. Sample paths of the model for a power-law adaptation current given by Eq. 3. Top: X(t) (Eq. 1), the horizontal dashed lines are at X = 0 and xth = 1. Bottom: s(t) (Eq. 2) When X reaches xth, s undergoes a jump of size κ. The subsequent ISIs Tk have distributions Fk(t), and the starting have distributions Gk(s) for k ≥ 1. Parameter values s(k) values are α = 3.0, γ = 1.0, σ = 0.8, I0 = 4.0, κ = 3.0. 0 In the non-renewal case we are studying here, subsequent firing times will in general not have the same distribution as T1. We define the kth interspike interval (ISI) as Tk = inf t − Ti : X(t) ≥ xth, t > Ti . (5) The first moment of the kth ISI is given by τ 1 k = E(Tk). The second moment of the kth ISI will be denoted by k = E((Tk)2) and the kth firing rate is given by the τ 2 inverse of the corresponding mean ISI rk = 1 . The kth τ 1 k standard deviation m2(k) is then given by m2(k) = k − (τ 1 τ 2 k )2 . (6) The values of the peak adaptation current after the kth firing are defined for k ≥ 1 as (cid:32) k−1(cid:88) i=1 (cid:33) k−1(cid:88) i=1 (cid:113) (cid:32) (cid:33) k(cid:88) i=1 ω(s) = − 1 τa s(t) , (4) s(k) 0 = s(t−) + κ : t = Ti , (7) 0.00.51.01.52.02.53.0−1.0−0.50.00.51.01.5X0.00.51.01.52.02.53.0t01234567ss(0)0s(1)0∼G1(s)T1∼F1(t)s(2)0∼G2(s)T2∼F2(t)s(3)0∼G3(s)T3∼F3(t) where t− indicates that we take the left-sided limit. For simplicity, we choose s to be started from a point (s(0) 0 = κ), instead of from a biophysically more realistic initial distribution. However, the methods we are going to describe in the following are also valid when s is ini- tially started from a distribution. The central challenge is to obtain the distributions Gk and Fk for k ≥ 1, which are the distributions of s(k) and Tk defined by Eqs. 7 and 5, respectively. An example realization for the case of power-law adaptation is shown in Fig. 1. The knowledge of these distributions is key to understanding the non-renewal dynamics, as they form a hidden Markov model of the underlying non-Markovian dynamics [15, 16, 32]. Therefore, once these distributions are known, the non-renewal dynamical system breaks up into coupled renewal dynamical systems, which are much more tractable mathematically. This gives rise to the iFPT approach which we now explain. 0 III. THE IFPT APPROACH Being a diffusion process, the system given by Eqs. 1 and 2 is governed by a two-dimensional time-dependent FPE [33]. The FPE for the probability density func- tion p(t; x, s)dxds = P(X(t) ∈ (x, x + dx), s(t) ∈ (s, s + ds)X(0) = x0, s(0) = s0) reads (we use x(cid:62) = (x, s) for brevity) ∂tp(t; x) = ∇ · (A(x)∇p(t; x)) − ∇ · (F(x)p( t; x)) , (8) where A is the diffusion matrix and F the drift vector, which can be obtained in a straightforward way from the SDE for X, Eq. 1, and the ODE for s, Eq. 2. Explicitly, we have (cid:18) φ(x)2 2 0 (cid:19) 0 0 , (9) A(x) = and F(x) = (µ(x) − s, ω(s)) (cid:62) . time T1 , CDF1(t) = (cid:82) t The IF property of X entails that we have an absorbing boundary at X = xth for all times t: p(t; x = xth, s) = 0. With this boundary condition, we can compute the cu- mulative distribution function (CDF) of the first-passage 0 F1(λ)dλ, by time evolution of the FPE on a computational domain Ω ⊂ R2, which we choose to be a rectangle extending to sufficiently negative values in the x-direction [34]: (cid:90) CDF1(t) = 1 − Ω p1(t; z)dz , (10) 3 0 0 , the value of s(k) where p1 is the solution to Eq. 8 with the initial condition p1(0; x, s) = δ(x − x0), where x0 = (0, κ)(cid:62). For the computation of F1, we then only need to differentiate Eq. 10. To describe adaptation, one needs to compute the statis- tics of the peak adaptation currents, as defined by Eq. 7. Hence, we need to characterize the hidden Markov model generated by the ISIs Tk and the peak adaptation cur- rents s(k) 0 . Whereas the dynamics of s and X as a whole is non-Markovian, the distribution of the ISI Tk is com- pletely determined by the distribution Gk−1, and the se- quence {Tk, s(k−1) ∞ k=1 is therefore Markovian. Knowing } the values of Tk and s(k−1) is fixed, and the distribution of the ISI Tk+1 can be obtained by solving an FPT problem with s(k) as initial condition for s. The central observation now is that for the second ISI, X again starts from 0, whereas s starts from a distribu- tion G1. This is because X evolves stochastically, and therefore reaches the threshold xth at different times T1, corresponding to different values of s(t = T1) + κ imme- diately after the firing event. To compute the second ISI, we therefore need to know G1. This can be iterated: to compute the distribution Fk of the kth ISI, we need the distribution Gk−1. This is the central idea of the iFPT approach. To set up the iFPT approach, we first observe that between threshold crossings of X, s evolves deter- ministically. Therefore, when we know the PDF of the first FPT, by conservation of probability, we also know the distribution of s after the first firing event: 0 0 (cid:12)(cid:12)(cid:12)(cid:12) dt(s) ds (cid:12)(cid:12)(cid:12)(cid:12)F1(t(s)) , G1 (s − κ) = (11) where t(s) is the inverse function of s. The support of G1 is shifted, because of the jump of size κ that s undergoes when X reaches its threshold xth. For the second ISI T2, s is started from the distribution G1(s) instead of a point, whereas X is started from a point again (Fig. 1). This means that to obtain F2, the FPE is started from a distribution: p2(0; x) ∝ G1(s)δ(x). This generalizes to values of k larger than 1. For the kth ISI distribution Fk(t), we therefore must choose pk(0; x) ∝ Gk−1(s)δ(x) . (12) We show how to obtain the distributions Gk for k > 1 in Section V. Linear splines are used to create a mesh be normalized appropriately, so that (cid:82) function approximating Eq. 12 on the computational do- main Ω. Due to this approximation, Eq. 12 then has to Ω pk(0; z)dz = 1. (cid:82) The FPE is then solved again, and the distributions Fk(t) are obtained analogously to Eq. 10: CDFk(t) = Ω pk(t; z)dz, i.e. by timestepping the FPE to obtain 1− the CDF of the kth ISI followed by a numerical differen- tiation. This constitutes the iFPT approach. To quantify the accuracy of our numerical methods, we also compute the relative disagreement ∆ between results obtained by the iFPT approach and direct MC simula- tions. It is defined for a quantity Z by ∆(Z) = ZiFPT − ZMC ZiFPT . (13) We performed MC simulations for two different simula- tion setups, the first one without any boundary correc- tion (plain MC), and the second one with a boundary correction according to Giraudo and Sacerdote (MC-GS) [31, 35, 36]. This boundary correction is applied to reduce the systematic overestimation of FPTs when using the Euler -- Maruyama scheme. We compute the relative dis- agreement given by Eq. 13 using either MC simulations with or without boundary correction; we observed that the order of the relative disagreement is unchanged, but in general, the plain MC algorithm gives rise to larger dis- agreements than MC-GS. A decrease in the relative dis- agreement is expected, because the GS correction method should yield an improved weak error of O(h) [37], in con- trast to the plain MC simulation, which has a weak error of O(h 1 2 ) [38], where h is the time step for the discretiza- tion of the SDE, Eq. 1. In the following, the timestep for MC simulations is chosen to be h = 10−3 and we choose M = 106 independent realizations. The plain Euler-Maruyama scheme then gives rise to a weak error 2 ) ≈ 3 · 10−2 when estimating moments of first of O(h 1 passage times, which is one order of magnitude larger = 10−3. There- than the MC error proportional to fore, in our simulations, the plain MC error is negligible in comparison to the error introduced by the finite-time discretization of the SDE (Eq. 1). For the numerical solution of the FPE (Eq. 8), we choose a finite element discretization method [39] and evolve the system using either a stabilized Crank -- Nicolson (CN) scheme [40] in Fig. 2 or an Euler timestepping scheme [39] in Fig. 3. The relative disagreement between MC simulations and finite-element solutions stays largely constant across dif- ferent lags k when we use the CN scheme instead of the Euler scheme as can be seen by comparing the lower pan- els of Figs. 2 and 3; the sizes of the disagreement are com- parable in magnitude. This suggests that the remaining small discrepancy between MC simulations and PDE re- sults can be largely explained with the errors associated with the MC simulation method. In particular, note that for the examples we show, the MC-GS weak error of size O(h) is comparable in magnitude to the numerator of ∆(Z) (Eq. 13), i.e. the absolute disagreement. We will see what effects this has on the computation of correla- tions in Section VI. 1√ M 4 FIG. 2. Top: Evolution of the rate rk = 1 (left) and stan- τ 1 k dard deviation given by Eq. 6 (right) of ISIs as a function of k. Empty circles: Plain MC simulations of Eqs. 1 and 2. Tri- angles: MC simulations with GS boundary correction. Filled circles: moments obtained from numerical solution of FPE using a CN timestepping scheme. M = 106 independent MC realizations for each value of k. Power-law adaptation (Eq. 3) with α = 5.5, I0 = 6.0, σ = 1.3, γ = 1.0, κ = 5.5. Bottom: Relative disagreements defined by Eq. 13, where an MC-GS algorithm was used. IV. TRANSITION TO STATIONARITY We show the evolution of the rate and standard deviation of ISIs in Figs. 2 and 3. The rates decreases, whereas the standard deviation increases until both quantities reach a stationary value. Given that these quantities are derived from moments of the ISI distributions, the distributions also converge towards a stationary form. The conver- gence towards stationarity of ISI and peak adaptation current distributions (Fk and Gk, respectively) is shown in Figs. 4 and 5. As expected for adapting models, the ISIs (whose distributions are shown in Fig. 4) increase with higher k, which means that the rate rk decreases. This is well captured by the iFPT approach, with a max- imal relative disagreement smaller than 3% in Fig. 2 and smaller than 2% in Fig. 3. Also, the width of both the ISI distributions and the peak adaptation current distribu- tions (Fig. 5) increases, which is reflected by the increase of the variances of Fk and Gk shown in Fig. 2 and 3. The mean of the peak adaptation currents shifts to the right as stationarity is reached. Moreover, stationarity is reached with varying speed, i.e. for different values of the lag k (compare Fig. 2 with Fig. 3). Generally, 123450.91.01.11.21.31.41.51.61.71.8rk123450.320.340.360.380.400.420.44m2(k)12345k0.0040.0060.0080.0100.0120.0140.0160.0180.0200.022∆(rk)12345k0.0040.0060.0080.0100.0120.0140.016∆(m2(k)) 5 FIG. 4. PDF Fk of the kth ISI Tk (Eq. 5). The symbols are MC simulations (M = 106 MC realizations) as indicated in the legends. Solid black lines: PDF obtained by numerical solution of the FPE, Eq. 8. In the left panel, the distributions are practically indistinguishable after k = 2. Left: Power-law adaptation (Eq. 3). Right: Single exponential adaptation (Eq. 4). Parameter values as in Fig. 2 for power-law adapta- tion and as in Fig. 3 for exponential adaptation. Both panels show results for plain MC simulations. value that is larger than those typically found around the mode of the stationary distribution) can build up. The system reaches stationarity rapidly because it is quasi- deterministic as the dynamics of s dominates the sys- tem, and the stochastic fluctuations of X will only cause small perturbations. For both power-law and single ex- ponential adaptation currents, it is possible to reach the stationary regime already after one or two firing events as in Fig. 2, or to have a long transient regime as in Fig. 3. The initial condition s(0) for the adaptation current can 0 also be chosen to control the speed of the transition. If it is placed far away from the mean of the stationary distri- bution, the transition will take a longer time; also, it is possible to obtain a non-monotonic behaviour of the rate as a function of the interval number when s(0) is placed 0 far above the aforementioned mean. The first mean ISI will then be the longest statistically, in contrast to the examples we show in Figs. 2 and 3. V. STATISTICS OF THE kTH ISI The iFPT approach can be iterated beyond the first two firing events to obtain the distribution for the third ISI, F3(t). However, for the computation of the third ISI, no equation similar to Eq. 11 can be used to obtain G2 because s was started from a distribution to obtain F2. Indeed, for one fixed time T2, there are many different starting values s(1) 0 due to the stochastic dynamics of X. Importantly, the ISI T2 and the initial condition s(1) are 0 not independent random variables (for a large value of s(1) 0 , a large ISI T2 is more probable and vice versa) so that we can obtain the value that s reaches after the second firing by the following observation (focusing on power-law adaptation): given that T2 = λ and s(1) 0 = ν, FIG. 3. Top: Evolution of the rate rk = 1 (left) and stan- τ 1 k dard deviation given by Eq. 6 (right) of ISIs as a function of k. Empty circles: Plain MC simulations of Eqs. 1 and 2. Tri- angles: MC simulations with GS boundary correction. Filled circles: moments obtained from numerical solution of FPE us- ing an Euler timestepping scheme. M = 106 independent MC realizations for each value of k. Single exponential adaptation (Eq. 4) with τa = 1.0, I0 = 5.0, σ = 1.0, γ = 1.0, κ = 1.0. Bottom: Relative disagreements defined by Eq. 13, where an MC-GS algorithm was used. the speed of adaptation can be controlled by adjusting the bias current I0 and the noise level σ as well as the adaptation strength (size of the kick κ and, in the case of single exponential adaptation, the timescale τa). We have carried out additional MC simulations (not shown) to obtain insight into how these parameters influence the speed of the transition to stationarity. A higher bias cur- rent and a higher noise level will in general lead to a less rapid transition to stationarity. This can be under- stood as follows: X is driven to threshold more rapidly, causing the inhibition to build up quickly, reaching val- ues that are higher than those typically found around the peak of the stationary distribution. This slows down the transition to stationarity, because s needs to decay first. Moreover, a large kick size κ and a rapidly de- caying adaptation current will cause a quick transition. For the latter case, this is easily understood as we are then nearly dealing with a renewal system: after a short initial period, the effect of the adaptation current on the firing time statistics is negligible. For the former case, we note that the larger the kick size κ, the more pronounced the inhibitory effect of adaptation within one ISI, which means that X takes longer to reach threshold before a large out-of-equilibrium average value of the current s (a 123456782.22.42.62.83.03.23.43.63.8rk123456780.120.140.160.180.200.220.24m2(k)12345678k0.0000.0020.0040.0060.0080.0100.0120.014∆(rk)12345678k0.0020.0040.0060.0080.0100.0120.0140.0160.0180.020∆(m2(k))0.00.51.01.52.02.53.0t10−210−1100Fk(t)k=1k=2k=30.00.20.40.60.8t10−210−1100k=1k=2k=3 6 0 FIG. 5. PDF Gk of s(k) (Eq. 7). The symbols are MC simula- tions (M = 106 MC realizations) as indicated in the legends. The black solid line is G1 obtained from Eq. 11 or G2 (G3) obtained using Eq. 18. In the left panel, the distributions are practically indistinguishable after k = 2. Left: Power-law adaptation (Eq. 3). Right: Single exponential adaptation (Eq. 4). Parameter values as in Fig. 2 for power-law adapta- tion and as in Fig. 3 for exponential adaptation. Both panels show results for plain MC simulations. FIG. 6. Conditional FPT densities H(λ, ν) given by Eq. 17 relating the initial value of the adaptation current to the dis- tribution of the following ISI. Computed using numerical so- lutions to the FPE Eq. 8. Left: Power-law adaptation, com- puted using a CN timestepping scheme. Right: Exponential adaptation, computed using an Euler timestepping scheme. Parameter values as in Fig. 2 for power-law adaptation and as in Fig. 3 for exponential adaptation. ν λ 0 = 1 α + 1 we have s(2) + κ. We have included the jump of size κ due to the definition of s(2) (see Fig. 1). This em- 0 phasizes that once two values in the triplet (T2, s(1) 0 , s(2) 0 ) are fixed, the third one is determined. In the following, we again use λ to denote a fixed FPT and ν to denote a fixed initial value of the adaptation current s. Anal- ogously, we generally have for s(k) 0 : given Tk = λ and s(k−1) 0 = f (λ, ν) is determined. The function f 0 determining the subsequent value of the peak adaptation current given the previous ISI λ and the previous peak value of the adaptation current ν reads for power-law adaptation (Eq. 3): = ν, s(k) f (λ, ν) = κ + 1 λ α + 1 ν , (14) whereas for exponential adaptation (Eq. 4), we have f (λ, ν) = κ + ν exp − . (15) An alternative way is to fix the value of s(k) and then put a constraint on the time Tk when s(k−1) is fixed: given s(k−1) 0 = θ, Tk = h(ν, θ) is deter- mined, where, for power-law adaptation, we have the ISI as a function of the previous and subsequent adaptation values: = ν and s(k) 0 0 0 (cid:18) (cid:19) λ τa (cid:18) 1 θ − κ − (cid:19) 1 ν h(ν, θ) = α . (16) The function h is defined by solving the equation f (λ, ν) = θ for λ. To actually compute the density Gk, (cid:16) (cid:17) we need to relate the above observations to densities that we can compute with the iFPT approach. To that end, we now define the conditional density H: H(λ, ν)dλ ≡ P T1 ∈ (λ, λ + dλ)s(0) 0 = ν . (17) We have used T1 and s(0) in the definition Eq. 17 to 0 stress that, for the purpose of the computation of H, we only need to solve the FPT problem for T1 using differ- ent values of the initial condition s(0) It will become 0 . apparent below that we only need to compute H once, because it does not depend on the firing index k. For a fixed value of ν, H(λ, ν) is an FPT probability den- sity. Our notation emphasizes that H is a function of two variables. ν sets the level of initial inhibition, i.e. the starting value of s. We show the function H for both power-law and exponential adaptation in Fig. 6. We see that with increasing starting value ν, the mode of the FPT distribution shifts to larger times. For power-law adaptation, the shape of the FPT distributions does not change much, whereas for exponential adaptation, the distributions become broader with increasing ν. With this at hand, we now show how to practically com- pute the distributions Gk for k > 1. We can obtain the CDF of s(k) 0 by observing that: (cid:17) (cid:90) D(k−1)(θ) H(λ, ν)Gk−1(ν)dλdν , s(k) 0 ≤ θ = (18) (cid:16) P with D(k−1)(θ) = (λ, ν > 0 ν ∈ supp (Gk−1) , h(ν, θ) ≤ λ ≤ Tmax) . The function h defined in Eq. 16 ensures that for a fixed value of ν, we collect all times λ so that s(k) 0 ≤ θ, which (19) 7891011s10−310−210−1Gk(s)k=1k=2k=31.52.02.53.03.5s10−210−1100k=1k=2k=3 ensures that f (λ, ν) ≤ θ for fixed values of θ and ν. Tmax is chosen so that H(Tmax, ν) ≈ 0 ∀ν ∈ supp (Gk−1). This means that Tmax should be chosen in the tail of the FPT distribution. Note that for the iFPT approach, one only has to compute H(λ, ν) over the support of Gk for k ≥ 1 once [41], and then multiply it with the adaptation cur- rent distribution of the previous iteration Gk−1. This function then needs to be integrated according to Eq. 18, and the PDF Gk can be obtained by numerical differ- entiation. We show results for G2 and G3 using Eq. 18 in Fig. 5. The agreement between MC simulations and Eq. 18 is excellent. Iterating these ideas into the stationary regime, the ideas of the previous paragraph can be used to directly com- pute the stationary density of the peak adaptation cur- rent. We assume that stationarity is reached after k∗ − 1 firing events. Then s(k∗) have the same dis- tribution. Consequently, the stationary density of the peak adaptation current after firing satisfies the two- dimensional integral equation (cf. Eq. 18) and s(k∗+1) 0 0 7 notion of stationarity, so that the SCC depends on both the position n of the ISI in the spike train as well as on the lag k between ISIs. Since we have already computed the distributions of the kth ISI, we can readily compute the variances and means in Eq. 21, i.e. the terms given by Eqs. 22 and 23. It is slightly more complicated to com- pute the first term in the numerator, E (TnTn+k), because we need the joint density p2(Tn, Tn+k) of Tn and Tn+k. In the present study, we focus on k = 1. By definition, we have E(TnTn+1) = dTndTn+1 Tn Tn+1 p2(Tn, Tn+1), = dTn dTn+1 Tn Tn+1p1(Tn+1Tn)Fn(Tn), where as previously Fn(Tn) is the density of the nth ISI and p1(Tn+1Tn) is the conditional density of Tn+1 given Tn. Because Tn+1 is statistically determined only by s(n) 0 , we can define this conditional density as Q(θ) = D(θ) H(λ, ν)Q (cid:48)(ν)dλdν , (20) p1(Tn+1Tn) = dy p(Tn+1, yTn), (cid:90) (cid:90) (cid:16) (cid:17) s(k∗) 0 ≤ θ denotes the CDF of the sta- where Q(θ) ≡ P tionary peak value for the adaptation current s (Eq. 2). We have checked that this equation is indeed satisfied by the stationary distributions for the peak adaptation current obtained from MC simulations (data not shown). Therefore, Eq. 20 can serve as a tool to check whether a given distribution for the peak adaptation current is stationary, or alternatively as a way to compute Q(θ) directly if the function H is known. VI. CORRELATIONS BETWEEN INTERSPIKE INTERVALS We now show how to compute serial correlations with the iFPT approach. We define the SCC [15, 42] between the nth ISI Tn and the (n + k)th ISI Tn+k according to where p(Tn+1, yTn) denotes the joint density of Tn+1 and s(n) 0 = y conditioned on the previous ISI Tn. We can rewrite this as follows: p(Tn+1, yTn) = p3(Tn+1, y, Tn) p(Tn) , = = p(Tn+1y, Tn)p(y, Tn) p(Tn) , p(Tn+1y, Tn)p(yTn)p(Tn) p(Tn) , = p(Tn+1y, Tn)p(yTn). Now, as we have previously shown, the statistics of Tn+1 is completely determined when s(n) 0 = y is fixed, hence p(Tn+1y, Tn) = p(Tn+1y) ≡ H(Tn+1, y). Therefore, we have (cid:90) (cid:90) (cid:90) (cid:90) SCC(n, k) = E (TnTn+k) − Q1(n, k) Q2(n, k) , (21) E(TnTn+1) = where (22) Q1(n, k) = E(Tn)E(Tn+k) , and Q2(n, k) = m2(k)m2(n + k) =(cid:112)Var(Tn)Var(Tn+k) . (23) Here, Var(Tn) denotes the variance of the nth ISI distri- bution, and m2 is the standard deviation given by Eq. 6. Note that the definition Eq. 21 does not make use of the dTn dTn+1 dy Tn Tn+1H(Tn+1, y)p(yTn)Fn(Tn) . This can be further simplified by noting that p(yTn) = Fn(Tn) and therefore p(y,Tn) E(TnTn+1) = dTn dTn+1 dy Tn Tn+1 H(Tn+1, y)p(y, Tn) . (24) For n = 1, Eq. 24 can be simplified because s(1) 0 deterministic function of T1 (see Eq. 11), so that p(s(1) is a 0 = This results in an alternative, approximative expression for the expectation E (TnTn+1): 8 (cid:90) E (TnTn+1) = dTndTn+1dy Tn Tn+1 H(Tn+1, y)Gn(y)Fn(Tn) . (27) Eq. 27 is therefore equivalent to Eq. 24 if p(y, Tn) = Gn(y)Fn(Tn). Although Fig. 7 demonstates that p does not factorise (the joint density is negatively sloped), we show below that Eq. 27 approximates Eq. 24 very well. Given that Eq. 27 does not require additional MC simu- lations, the small error introduced by Eq. 27 is well offset by the large reduction in computational cost. There exists a third alternative expression for the expec- tation of the product of ISIs suggested for a different, but related, model, in [43]. It reads in our notation [44] (cid:90) E (TnTn+1) = dTn dTn+1 dy Tn Tn+1 H(Tn+1, f (Tn, y))H(Tn, y)Gn−1(y) , (28) where f is given by Eq. 14 or Eq. 15. The term H(Tn, y)Gn−1(y) that appears in Eq. 28 is the same as the one on the right-hand side in Eq. 18, which we used to obtain the CDF of s(n) 0 . For n = 1, Eq. 28 reads (cid:90) E (T1T2) = dT1 dT2 T1 T2 H (cid:16) T2, f (cid:16) T1, s(0) 0 (cid:17)(cid:17) (cid:124) H (cid:16) (cid:123)(cid:122) T1, s(0) 0 =F1(T1) (cid:17) (cid:125) , (cid:16) (cid:17) 0 y − s(0) because s is started from a point s(0) 0 , so that formally , which collapses the integration over G0(y) = δ y in Eq. 28. Thus, for n = 1, Eq. 24 and 28 coincide. This is also true for higher values of n. A proof for this equiva- lence is presented in Appendix A. Eq. 28 only makes use of the quantities H and G, which can be computed using the iFPT approach as explained in the previous section. We show comparisons between MC simulations and the three expressions for E (TnTn+1), Eq. 24, Eq. 27 and Eq. 28, in Fig. 8. The results presented in Fig. 8 are in agreement with the observation that the two expres- sions Eq. 24 and Eq. 28 are equivalent. We find that the agreement of Eq. 28 with MC simulations is comparable to Eq. 27, particularly for exponential adaptation. Inter- estingly, the formally correct Eq. 24 and the approximate Eq. 27 give comparable results; Eq. 24 slightly deviates from MC simulations and Eq. 27 when n gets larger. The maximal relative disagreement between MC and iFPT re- sults is less than 2% (Fig. 8, bottom panels). We will see below that the SCC is best approximated by using the ex- act result Eq. 24 (or equivalently Eq. 28), as we expect. 0 FIG. 7. MC simulations with GS boundary correction for the joint density of s(n) (i.e. nth peak value of adaptation current following the nth FPT) and Tn for different values of n. Left: Power-law adaptation, n = 2. Right: Exponential adaptation, n = 3. M = 106 MC realizations. Parameter values as in Fig. 2 for power-law adaptation and as in Fig. 3 for exponential adaptation. y, T1 = x) = δ by Eqs. 14 and 15. Hence, we have for n = 1 y − f (x, s(0) 0 ) F1(x), where f is defined (cid:17) (cid:16) (cid:90) E(T1T2) = dT1 dT2 T1 T2 H(T2, f (T1, s(0) 0 ))F1(T1). 0 (cid:17) s(n) 0 = y, Tn We have shown in Section V how to obtain the con- (cid:16) ditional FPT density H. To evaluate Eq. 24 for gen- eral n, we still need to compute the joint density . This can be achieved by means of an p MC simulation, where we fix a value of n and then record the frequency with which pairs of s(n) and Tn are gener- ated by the system. We show an example of these den- sities in Fig. 7. The most notable feature is an inverse proportionality between s(n) and Tn. The longer e.g. T2, the less likely it is for the value of s after the second firing, s(2) Eq. 24 is formally correct, but not very practical for ac- tual computations. This is because to apply the iFPT approach, it is desirable to obtain all quantities needed for the SCC using solutions of the FPE only, and no MC simulations. These, however, are required to ob- tain an approximation for the joint density p(s(n) 0 , Tn) in Eq. 24. We therefore propose an approximation to com- pute E (TnTn+1) using the available densities F, G and H only. To that end, we note that 0 , to attain a high value. 0 p3(Tn+1, Tn, y) = p(Tn+1, Tny)p(y) . (25) If we now assume that Tn and s(n) can approximate this as follows: 0 are independent, we p3(Tn+1, Tn, y) ≈ p(Tn+1y)p(y)p(Tn) = H(Tn+1, y)Gn(y)Fn(Tn) . (26) We attribute the discrepancy between MC simulations and the exact result Eq. 24 to the error caused by the numerical integration over the MC approximation of the joint density p(y, Tn). We checked that applying a kernel density estimation [45] to the MC results for p(y, Tn) did not alter these results. Similar results for Q1 and Q2 (Eqs. 22 and 23) are shown in Figs. 9 and 10. The agreement is good, with the maxi- mal relative disagreement always less than 5%. The rela- tive disagreement for the statistics of the product of two adjacent ISIs, Q1(n, 1), is in general larger than the error for the moments, as can be seen by comparing Figs. 2 and 3 with Fig. 9. Indeed, for the case of power-law adapta- tion, we observe an increase of roughly one order of mag- nitude in the relative error even when the more accurate CN scheme is used (see e.g. left panels of Fig. 9). An exception is the computation of the joint expectation, shown in Fig. 8, where, depending on which methods are compared, the relative disagreement is comparable in size to the one for the computation of the moments shown in Figs. 2 and 3. This increase of the relative dis- agreement makes the computation of the SCC using the iFPT approach hard, because the two expressions in the numerator of Eq. 21 are quite close to one another for the parameter values we have chosen here, meaning that the numerator is small and indeed of the same magnitude or even smaller as the relative disagreement, e.g. −2.6·10−2 in the left panel and −8.6 · 10−4 in the right panel of Fig. 8 for n = 1 for the MC-GS simulation method. The increase in the relative discrepancy is caused by error propagation, because for the second-order statistics, one has to multiply two quantities that both come with an individual error. Therefore, whereas the iFPT approach can in principle also be used to compute serial correla- tions present in the spike train, obtaining reliable results can in general be a computational challenge. When the negative serial correlations are stronger, so that the dif- ference in the numerator of Eq. 21 is larger, the iFPT approach should give more accurate results. We stress that the dominant source of error is not the computation of the joint expectation E (TnTn+1) of ISIs, but the prod- uct of the expectation of ISIs and the variances, which can be seen by comparing the lower panels of Fig. 8 with those of Figs. 9 and 10. Finally, we show the SCC at lag 1 obtained by MC simu- lations and PDE numerics in Fig. 11. The agreement is worse than for all previously considered quantities, but still reasonable. To verify the MC simulations, we checked that our MC simulation setup was able to re- produce known analytical results for the SCC obtained in [15] for certain limiting cases. The anti-correlations between adjacent ISIs (SCC(n, 1) < 0) strengthen until they reach a stationary value. Thus, we see that MC and PDE results for the SCC in general do not agree as well as one would expect from the good agreement of the expecations E(TnTn+1) in Fig. 8. The deviation is likely more pronounced for parameters that lead to small negative SCCs, which we have for both 9 FIG. 8. Top: Expectation E (TnTn+1) of adjacent ISIs for different values of n. Left: power-law adaptation. Right: exponential adaptation. M = 106 MC realizations. Empty circles: Plain MC simulations of Eqs. 1 and 2. Triangles: MC simulations with GS boundary correction. Pentagons: Eq. 24. For power-law adaptation, the pentagons are on top of the empty triangles and filled circles and hence not visi- ble. Filled circles: Eq. 27. Diamonds: Eq. 28. The diamonds are nearly on top of the filled circles and hence not visible. The PDE results were obtained using a CN scheme (power- law adaptation) or an Euler timestepping scheme (exponential adaptation). The vertical error bars show the MC error for a > 99.99% confidence interval. Bottom: Relative disagree- ments defined by Eq. 13, where for power-law adapation, the GS boundary corrected MC algorithm was used, and for ex- ponential adaptation, the plain MC algorithm was used. For the iFPT quantity, Eq. 24 was used. Parameter values as in Fig. 2 for power-law adaptation and as in Fig. 3 for exponen- tial adaptation. models considered in this section. In the next Section, we will compare this with results for the perfect integrate- and-fire model, where parameter values are chosen so that the SCCs are more negative and hence the agree- ment is better. This is because the two terms in the nu- merator of Eq. 21 are close to each other for small SCCs, and hence a small error in them impacts the accuracy of the SCC computation quite dramatically. We show in the next section that our methods reproduce known stationary analytical results for the SCC when we consider the perfect integrate-and-fire model with single exponential adaptation in a parameter regime where we have large negative correlations, thus demonstrating that our methods are sound, but SCC calculations are very sensitive to numerical inaccuracies. 123450.50.60.70.80.91.0E(TnTn+1)12345n0.0000.0020.0040.0060.0080.0100.0120.014∆(E(TnTn+1))12345670.080.100.120.140.160.181234567n0.0020.0040.0060.0080.0100.0120.0140.016 10 FIG. 9. Top: Q1(n, 1) (defined in Eq. 22) of adjacent ISIs for different values of n. Left: power-law adaptation. Right: exponential adaptation. Empty circles: MC simulations of Eqs. 1 and 2. Triangles: MC simulations with GS boundary correction. Filled circles: Eq. 22, CN timestepping scheme for power-law adaptation and Euler timestepping scheme for exponential adaptation. Bottom: Relative disagreement de- fined by Eq. 13, a plain MC algorithm was used to obtain the relative disagreement. Parameter values as in Fig. 2 for power-law adaptation and as in Fig. 3 for exponential adap- tation. FIG. 10. Top: Q2(n, 1) (defined in Eq. 23) for adjacent ISIs for different values of n. Left: power-law adaptation. Right: exponential adaptation. Empty circles: MC simulations of Eqs. 1 and 2. Triangles: MC simulations with GS boundary correction. Filled circles: Eq. 23, CN timestepping scheme for power-law adaptation and Euler timestepping scheme for exponential adaptation. Bottom: Relative disagreement de- fined by Eq. 13, a plain MC algorithm was used to obtain the relative disagreement. Parameter values as in Fig. 2 for power-law adaptation and as in Fig. 3 for exponential adapta- tion. Parameter values as in Fig. 2 for power-law adaptation and as in Fig. 3 for exponential adaptation. A. The perfect integrate-and-fire model The adapting perfect integrate-and-fire (PIF) model driven by white Gaussian noise and a single exponen- tial adaptation current is one of the simplest models for spike-triggered adaptation. For small noise intensity, an- alytical expressions for the stationary SCC exist. We here study this stationary limit case and compare analytical formulas to results obtained with the iFPT approach. The model reads (we follow the notation of [46] and [47]) dX = (I0 − s)dt + √2DdW (t) , ds dt = − s τa . (29) (30) The adaptation mechanism works in analogy to the pre- vious model (Eq. 2): whenever X reaches the threshold X = 1, s receives a kick of size ∆ ≡ and X is instan- taneously reset to 0. The stationary SCC at lag 1 for this model under the assumption of small noise (i.e. D (cid:28) 1) reads [46] (cid:101)∆ τa SCC(k = 1) = − α(1 − θ)(1 − α2θ) 1 + α2 − 2α2θ , (31) where α = T ∗ = s∗ − ∆ s∗ 1 +(cid:101)∆ I0 , θ = , s∗ = , I0 − s∗ I0 − s∗ + ∆ (cid:16) − T ∗ ∆ 1 − exp τa (cid:17) . Thus, we can compute the SCC in closed analytical form as a function of the system parameters. This formula serves as an important benchmark for our numerical re- sults. In particular, we expect that after the described transition to stationarity, the SCC given by Eq. 21 will approach the stationary SCC given by Eq. 31. This is confirmed in Fig. 12. In particular, the agreement be- tween MC simulations and the exact formula Eq. 24 is very good (the relative disagreement between PDE nu- merics and the analytical result is less than 6% for the 12340.50.60.70.80.91.01.1Q1(n,1)1234n0.0180.0190.0200.0210.0220.0230.0240.0250.026∆(Q1(n,1))12345670.080.100.120.140.160.180.201234567n0.0370.0380.0390.0400.0410.0420.0430.0440.04512340.1400.1450.1500.1550.1600.1650.1700.1750.1800.185Q2(n,1)1234n0.0020.0040.0060.0080.0100.0120.014∆(Q2(n,1))12345670.0200.0250.0300.0350.0400.0450.0501234567n0.0000.0050.0100.0150.0200.025 11 FIG. 11. Serial correlation coefficient at lag n = 1 (defined in Eq. 21) for adjacent ISIs for different values of n for the LIF model (Eq. 1). Left: power-law adaptation. Right: expo- nential adaptation. Empty circles: Plain MC simulations of Eqs. 1 and 2. Triangles: MC simulations with GS boundary correction. Filled circles: PDE results, SCC computed using Eq. 27. Pentagons: PDE results, SCC computed using Eq. 24. CN timestepping scheme for power-law adaptation and Euler timestepping scheme for exponential adaptation. Diamonds: PDE results, SCC computed using Eq. 28. Parameter val- ues as in Fig. 2 for power-law adaptation and as in Fig. 3 for exponential adaptation. Even if the joint expectations shown in Fig. 8 agree well, this does not imply that the SCC will be well approximated; the small correlation values (i.e. the difference between E(TnTn+1) and Q1(n, 1)) for the two examples lead to large discrepancies in the SCCs, which are especially significant for the case of power-law adaptation. stationary value); the agreement of MC simulations with the approximation Eq. 27 is a bit worse, but still rea- sonable. Thus, we conclude that our methodology can be used more generally to compute the evolution of the moments and SCCs. However, as seen in the previous section, to obtain a good agreement between MC simula- tions and PDE numerics, the computational effort might be rather large. In particular, we note that the PIF ex- ample shown in Fig. 12 gives rise to stronger negative SCCs, which means that the error propagation has less of an effect, but is still present, even when moments of firing times between MC and PDE numerics disagree by less than 1% (data not shown). We finally note that it is also possible to analytically compute the stationary SCC at higher lags and for different models (e.g. the leaky integrate-and-fire model in the presence of weak noise or for small adaptation currents) using the approach de- scribed in [47], or, using a different approach, in [15]. FIG. 12. Serial correlation coefficient at lag n = 1 (defined in Eq. 21) for different values of n for the PIF model (Eq. 30). The dashed horizontal line is the stationary SCC given by Eq. 31. Empty circles: Plain MC simulations of Eq. 30. Pen- tagons: Eq. 24. Filled circles: Eq. 27. The PDE results were obtained using a CN timestepping scheme. M = 106 MC realizations. Timestep h = 10−3. Parameter values: D = 0.1, τa = 5.0, (cid:101)∆ = 10, I0 = 5.5, s(0) 0 = 5.0. VII. SUMMARY AND CONCLUSIONS In this paper, we have developed a numerical method for the computation of moments and correlations in gen- eral two-dimensional non-renewal escape time processes. Our approach relies on the numerical solution of a two- dimensional time-dependent FPE with initial conditions obtained from marginal distributions of previous states of the system. Crucially, the computation scheme presented in this study is general insofar as it can be applied to any stochastic process with a known reset condition (Eq. 1) and any deterministic signal (Eq. 2). As an important application, we have described the transition to station- arity in a stochastic IF neuron model with spike-triggered adaptation, which causes non-trivial ISI correlations. A different mechanism for introducing positive correlations between ISIs has recently been reported in [28] and can equally well be analyzed with the presented methodol- ogy. Moreover, our approach enables us to determine the non-trivial timescale of transition to a stationary adapted state by counting the number of intervals needed for this transition. Experimentally, the transition to stationarity is often characterized by the behaviour of the instantaneous firing rate [17, 18] [48]. The instantaneous firing rate is usually obtained by averaging the neuronal activity bin-wise for a fixed time. This differs from the firing rate used here as given by the inverse of the mean ISI (Eq. 5). In other words, while the instantaneous firing rate is measured in real time, our firing rate relates to interval numbers. This entails that for a given time t, the firing rate contains con- tributions from, in general, past firing events that may have occurred at any point k in the spike train. Knowing the joint distributions of all ISIs Tk, it is at least in princi- ple possible to reconstruct the instantaneous firing rate, whereas given the instantaneous firing rate, we cannot 1234n−0.35−0.30−0.25−0.20SCC(n,1)1234567n−0.20−0.15−0.10−0.050.000.050.100.15123n−0.80−0.75−0.70−0.65−0.60−0.55−0.50SCC(n,1) reconstruct the joint distributions of the individual ISIs Tk. Despite the difference in the definition of the firing rate, it might be an interesting topic for further study to classify the time scales of the transition to stationarity both experimentally and based on the theory presented here. The computation of ISI moments using the iFPT ap- proach is computationally inexpensive, giving rise to small relative disagreements between solutions of the FPE and direct MC simulations. In contrast, the com- putation of correlations is harder. We observed that we lost one order of magnitude in accuracy compared to the simulation of the moments for the quantities Q1 (Eq. 22) and Q2 (Eq. 23), which makes the reliable computation of SCCs a computationally challenging task. We con- clude that even a relative disagreement of ISI moments between Monte Carlo estimations and PDE solutions of the order 10−3 is not enough to reliably estimate the SCC using PDE numerics only (but this might be spe- cific for the examples we have considered), indicating that more refined numerical methods or larger computational ressources, or indeed both, are needed. When the differ- ence between the two terms in the numerator of Eq. 21 is large, the small error made by the numerical solution of the PDE should have a less detrimental influence on the final result. The need for more refined numerical meth- ods is further substantiated by the fact that the more accurate asymptotically stable CN timestepping scheme did not result in a significant decrease in the relative dis- agreement between PDE results and both plain MC and MC-GS simulations, for both moments of firing intervals and the SCC. In this paper, we have only discussed the error associated with MC simulations, because it is read- ily available. The numerical solution of the FPE is of course also subject to numerical errors and future work will likely benefit from a discussion about how to sys- tematically reduce these errors. In this context, it might be beneficial to compare the finite-element methods used here to other methods for solving PDEs, such as finite difference and finite volume methods [49]. A systematic error estimation study might be made more difficult by the fact that the diffusion matrix (Eq. 9) is not positive definite [50, 51]. There is an alternative method to compute the ISI dis- tributions given the distributions of the peak adaptation currents using the formula (cid:90) supp(Gk−1) H(t, y)Gk−1(y)dy . Fk(t) = This is an integral equation frequently used in the con- text of randomized FPT problems [52, 53], where usu- ally Fk and the kernel H are given, and one tries to find a matching distribution Gk−1 of starting points. Using Eq. 32, we do not have to solve a time-dependent PDE for each ISI, but must compute H once as the solution of a time-dependent FPE with varying initial conditions for s, similar to the computation of F1. The averaging (32) E(TnTn+1) = (A1) dTn dTn+1 ds(n) 0 Tn Tn+1 H(Tn+1, s(n) 0 )p(s(n) 0 , Tn) . (cid:90) (cid:90) 12 that the iFPT approach amounts to is particularly clear in this formulation. The densities Gk are obtained as discussed above (see Eq. 18). The approach relying on Eq. 32 might be computationally less expensive, but we found that it is not as exact as solving a time-dependent FPE for each ISI, especially at larger times. This is likely caused by errors when computing H, as the numerical integration in Eq. 32 can be performed accurately and efficiently. However, Eq. 32 could be useful for analytical explorations when H is known. We finally emphasize that our approach did not use the complicated boundary conditions for stationary IF mod- els, where the probability flux at threshold gives rise to a discontinuity of the probability flux at reset [9, 54]. In contrast, our approach allows for the computation of transient and stationary distributions of the adapta- tion dynamics in an iterative fashion, requiring the solu- tion of a two-dimensional time-dependent PDE. The only boundary condition that has to be taken into account is an absorbing boundary condition for the probability density at the threshold xth. This makes the problem tractable using finite-element approximation methods for time-dependent PDEs, resulting in a general description of two-dimensional IF models with spike-triggered adap- tation. The approach we have described in this paper can in principle also be used to gain analytical insight into these system, however, quantities such as H and the solution of a two-dimensional time-dependent PDE seem to be unavailable in closed analytical form except in the most simple cases. ACKNOWLEDGMENTS We would like to thank Alexandre Payeur for insightful discussions and helpful comments (especially about the equivalence of Eqs. 24 and 28) and NSERC Canada for funding this work. Furthermore, we would like to thank an anonymous referee for comments that helped us to improve the manuscript. Appendix A: Equivalence of Eqs. 24 and 28 We here show that Eq. 24 and Eq. 28 are equivalent. We recall Eq. 24: We re-write Eq. 28 as follows: E (TnTn+1) = dTn dTn+1 dy Tn Tn+1 H (Tn+1, f (Tn, y)) p(y, Tn) , (A2) s(n−1) 0 y = have and we where replaced H(Tn, y)Gn−1(y) = p(y, Tn). Note that in Eq. A1, p is the joint density of s(n) and Tn, whereas p is the joint density of s(n−1) By inspection, the two expressions are identical if we can show that and Tn in Eq. A2. 0 0 13 0 to s(n) ), we have 0 = f (Tn, s(n−1) for Tn and Tn+1 fixed. Starting from the second line in Eq. A3, we change the integration variable from s(n−1) 0 by observing that ds(n) from s(n) and ds(n−1) therefore ds(n−1) . We need to as- sume that f is invertible with respect to the second ar- gument, which is the case for both power-law (Eq. 14) and exponential adaptation (Eq. 15) considered in this paper. The integral then becomes (cid:19)−1 = ds(n) = ∂f ∂s(n−1) ∂s(n−1) (cid:18) ∂f 0 0 0 0 0 0 0 (cid:90) (cid:90) (cid:104) 0 )p(s(n) 0 , Tn) = H(Tn+1, f (Tn, s(n−1) 0 ))p(s(n−1) 0 , Tn) , (A3) ds(n) ds(n−1) 0 H(Tn+1, s(n) (cid:90) 0 dTn dTn+1 ds(n) 0 Tn Tn+1 H (cid:16) (cid:17) (cid:104) p Tn+1, s(n) 0 ) f−1(Tn, s(n) 0 ), Tn (cid:105)(cid:32) (cid:33)−1 . ∂f ∂s(n−1) 0 (cid:105)(cid:18) (cid:19)−1 is nothing but so that , Tn) to p(s(n) 0 , Tn). In- f−1(Tn, s(n) 0 ), Tn) But p ∂s(n−1) the transformation from p(s(n−1) deed, we have (fixing Tn) ∂f 0 0 (cid:104) f−1(Tn, s(n) 0 ), Tn (cid:105)(cid:32) ∂f ∂s(n−1) 0 p(s(n) 0 , Tn) = p (A4) (cid:33)−1 . (A6) p(s(n) 0 , Tn) 0 ∂s(n) ∂s(n−1) 0 = p(s(n−1) 0 , Tn) , (A5) Therefore, Eq. 24 and Eq. 28 are equivalent. [1] A. Palacios, J. Aven, P. Longhini, V. In, and A. R. Bul- J. Benda, J Neurosci. 32, 17332 (2012). sara, Phys. Rev. E 74, 021122 (2006). [2] A. Aragoneses, L. Carpi, N. Tarasov, D. V. Churkin, M. C. Torrent, C. Masoller, and S. K. Turitsyn, Phys. Rev. Lett. 116, 033902 (2016). [3] M. J. Chacron, B. Lindner, and A. Longtin, Phys. Rev. Lett. 92 (2004). [4] J. A. Reinoso, M. C. Torrent, and C. Masoller, Phys. Rev. E 94, 032218 (2016). [5] S. Coombes, R. Thul, J. Laudanski, A. R. Palmer, and C. J. Sumner, Frontiers in Life Science 5, 91 (2011). [6] W. H. Nesse, L. Maler, and A. Longtin, Proc. Nat. Acad. Sci. USA 107, 21973 (2010). [7] D. R. Cox and P. A. W. Lewis, The Statistical Analysis of Series of Events, Methuen's Monographs on Applied Probability and Statistics (John Wiley, London, 1966). [8] R. Rosenbaum, Front. Comput. Neurosci. 10 (2016). [9] T. Schwalger and B. Lindner, Phys. Rev. E 92, 062703 (2015). [15] E. Urdapilleta, Phys. Rev. E 84, 041904 (2011). [16] E. Urdapilleta, Europhysics Letters 115 (2016). [17] R. Naud and W. Gerstner, PLOS Computational Biology 8, 1 (2012). [18] G. La Camera, A. Rauch, D. Thurbon, H.-R. Lüscher, W. Senn, and S. Fusi, Journal of Neurophysiology 96, 3448 (2006). [19] N. Ulanovsky, L. Las, D. Farkas, and I. Nelken, Journal of Neuroscience 24, 10440 (2004). [20] C. Kuehn, Multiple Time Scale Dynamics, 1st ed., Ap- plied Mathematical Sciences No. 191 (Springer-Verlag, 2015). [21] P. J. Drew and L. Abbott, J Neurophysiol 96, 826 (2006). [22] B. N. Lundstrom, M. H. Higgs, W. J. Spain, and A. L. Fairhall, Nat. Neurosci. 11, 1335 (2008). [23] C. Pozzorini, R. Naud, S. Mensi, and W. Gerstner, Nat. Neurosci. 16, 942 (2013). [24] S. E. Clarke, R. Naud, A. Longtin, and L. Maler, Proc. [10] Y.-H. Liu and X.-J. Wang, Journal of Computational Nat. Acad. Sci. USA 110, 13624 (2013). Neuroscience 10, 25 (2001). [25] R. Kobayashi, Y. Tsubo, and S. Shinomoto, Front. Com- [11] E. Muller, L. Buesing, J. Schemmel, and K. Meier, Neu- put. Neurosci. 3, 121 (2009). ral Computation 19, 2958 (2007). [12] L. Hertäg, D. Durstewitz, and N. Brunel, Frontiers in Computational Neuroscience 8 (2014), 10.3389/fn- com.2014.00116. [13] T. Schwalger, K. Fisch, J. Benda, and B. Lindner, PLoS Comput. Biol. 6 (2010). [26] W. Gerstner and R. Naud, Science 326, 379 (2009). [27] S. Redner, A Guide to First-Passage Processes, 1st ed. (Cambridge University Press, 2010). [28] G. D'Onofrio, E. Pirozzi, and M. Magnasco, in EU- ROCAST 2015, LNCS 9520, edited by R. M.-D. et al. (Springer-Verlag, Berlin Heidelberg, 2015) pp. 166 -- 173. [14] K. Fisch, T. Schwalger, B. Lindner, A. Herz, and [29] A. N. Burkitt, Biol. Cybern. 95, 1 (2006). [30] J. Benda and A. V. M. Herz, Neural Computation 15, (2016). 14 [43] M. J. Chacron, K. Pakdaman, and A. Longtin, Neural [44] Note Computation 15, 253 (2003). 3.17 (cid:82) dTndTn+1dy TnTn+1H(Tn+1, f (Tn, y))H (Tn, y)G∗(y). E (TnTn+1) stationary state: that Eq. the in [43] is in = estimation," [45] "Scipy v0.18.1 reference guide for Gaussian kernel http://docs.scipy.org/doc/ density scipy-0.18.1/reference/generated/scipy.stats. gaussian_kde.html, accessed: 2017-01-31. interspike-interval statistics of non-renewal neuron models, Ph.D. thesis, Humboldt- Universität zu Berlin (2013). [46] T. Schwalger, The [47] T. Schwalger and B. Lindner, Frontiers in Computational Neuroscience 7 (2013). [48] Note that in [18], the timescale for single exponential adaptation was inferred from the time course of the numerically obtained instantaneous firing rate, showing that the time course of the rate is not described by the same single exponential time course of the adaptation current. A similar observation is made in [30]. [49] R. Rosenbaum, F. Marpeau, J. Ma, A. Barua, and K. Josić, Journal of Mathematical Biology 65, 1 (2012). [50] S. Salsa, Partial Differential Equations in Action, 1st ed. (Springer-Verlag, 2008). [51] A. Ern and J.-L. Guermond, Theory and Practice of Fi- nite Elements, 1st ed., Applied Mathematical Sciences, Vol. 159 (Springer-Verlag New York, 2004) p. 526. [52] K. Jackson, A. Kreinin, and W. Zhang, Statistics and Probability Letters 79, 2422 (2009). [53] S. Jaimungal, A. Kreinin, and A. Valov, SIAM Theory Probabl. Appl. 58, 493 (2014). [54] N. Brunel and S. Sergi, J. Theor. Biol. 195, 87 (1998). 2523 (2003). [31] L. Sacerdote and M. T. Giraudo, in Stochastic Biomath- ematical Models, (M.Bachar et al. (eds.)), Lecture Notes in Mathematics 2058, Chapter 5. , 99 (2013). [32] C. van Vreeswijk, in Analysis of Parallel Spike Trains, Springer Series in Computational Neuroscience, Vol. 7, edited by S. Grün and S. Rotter (Springer Verlag, 2010) pp. 3 -- 20. [33] H. Risken, The Fokker-Planck equation: Methods of so- lution and applications, 2nd ed. (Springer, 1989). [34] B. F. Spencer Jr. and L. Bergman, Nonlinear Dynamics 4, 357 (1993). [35] M. T. Giraudo and L. Sacerdote, Communications in Statistics - Simulation and Computation 28, 1135 (1999). [36] B. Lindner and A. Longtin, Journal of Theoretical Biol- ogy 232, 505 (2005). 87, 167 (2000). [37] E. Gobet, Stochastic Processes and their Applications [38] D. Higham, X. Mao, M. Roj, Q. Song, and G. Yin, SIAM/ASA Journal on Uncertainty Quantification 1, 2 (2013). [39] A. Logg, K.-A. Mardal, and G. Wells, eds., Automated Solution of Differential Equations by the Finite Element Method, Lecture Notes in Computational Science and En- gineering, Vol. 84 (Springer Verlag, 2012). [40] P. Knabner and L. Angermann, Numerical methods for elliptic and parabolic partial differential equations, Texts in Applied Mathematics, Vol. 44 (Springer Verlag, New York Berlin Heidelberg, 2003). 0 + k · κ). [41] The support of Gk is the open interval (0, s(0) [42] R. Mankin and A. Rekker, Phys. Rev. E 94, 062103
1603.00397
1
1603
2016-02-29T18:44:23
A Langevin model for complex cardiological time series
[ "q-bio.NC", "cond-mat.stat-mech" ]
There has been considerable efforts to understand the underlying complex dynamics in physiological time series. Methods originated from statistical physics revealed a non-Gaussian statistics and long range correlations in those signals. This suggests that the regulatory system operates out of equilibrium. Herein the complex fluctuations in blood pressure time series were successful described by physiological motivated Langevin equation under a sigmoid restoring force with multiplicative noise.
q-bio.NC
q-bio
A Langevin model for complex cardiological time series C.E.C. Galhardo1, B. C. Coutinho2, T.J.P.Penna3, M.A. de Menezes3 and P.P.S. Soares4 1 Materials Metrology Division, National Institute of Metrology, Quality and Technology, Duque de Caxias, RJ ,Brazil 2 Petrobr´as, Cenpes, Rio de Janeiro, RJ, Brazil 3 Instituto de F´ısica, Universidade Federal Fluminense, Niteroi, RJ ,Brazil and 4 Instituto Biom´edico, Universidade Federal Fluminense, Niteroi, RJ, Brazil There has been considerable efforts to understand the underlying complex dynamics in physiolog- ical time series. Methods originated from statistical physics revealed a non-Gaussian statistics and long range correlations in those signals. This suggests that the regulatory system operates out of equilibrium. Herein the complex fluctuations in blood pressure time series were successful described by physiological motivated Langevin equation under a sigmoid restoring force with multiplicative noise. PACS numbers: 05.40.-a,87.80.Vt,87.19.L-,87.19.Hh Introduction - The autonomic nervous system is able to maintain life signals at safe levels by the action of a pair of nerve branches, called sympathetic and parasym- pathetic. While the sympathetic prepares our body for "flight or fight" the parasympathetic (or vagal) is con- sidered as a "rest and digest" system. In many cases, to achieve the homeostatic optimal levels, these systems have a competitive approach: while one start up an phys- iological reaction the other one suppress it [1]. Homeostasis depends on the blood flow according to the metabolic demands of each body part. The exchange of nutrients and metabolites occurs when blood flows through capillary channels. The perfusion either into or out the capillary depends on the blood pressure. Ade- quate levels of blood pressure are controlled by several mechanisms that can be classified according to the re- sponse delay: the long term and short term control. In order to maintain homeostasis the body automat- ically responds to changes. These responses are called reflexes. The principal short term reflex regulation of ar- terial blood pressure is the baroreflex. Stretch-sensitive mechanoreceptors are located in the carotid sinus and aortic arch connected to the brainstem, or nucleus trac- tus solitarii, by the glossopharyngeal and vagal nerves. After integrating the afferent signals the central nervous system, in turn, excite/inhibit the vagal branch if the pressure is high/low enough, closing the circuit for what can be regarded as a self-inhibitory feedback [2]. In addition, blood pressure may vary to adapt different physiological conditions such as exercise [3] or pregnancy [4] and in certain disease states such as hypertension [5]. The optimal level of blood pressure must be risk adjusted: cannot be high enough to cause structural damage and cannot be low enough to hinder the nutrient flow. It has been suggest that the underlying cardiac con- trol system can be characterized as a complex system. Indeed, heart and blood pressure present fractal time se- ries with long range correlations and non-Gaussian dis- tributions [6] and undergoes a breakdown of critical char- acteristics like a continous second ordem phase transtion [7]. In this letter, we propose an analytically tractable stochastic model for the baroreflex that capture the frac- tality and the non-Gaussian behavior in the blood pres- sure time series. This model leads to solutions in terms of the so called q-Gaussians, which is a well known function in the framework of Nonextensive Statistical Mechanics. This theory have shown excellent results describing time series systems with fractal structure and an introduction and many examples of the theory can be found in [8]. Experiment - Two groups of adult male Wistar rats were analyzed: control rats and chronic sinoaortic dener- vated rats, animals surgically denervated 20 days before measurements. Sinoaortic denervated was performed us- ing the methods described by Krieger et al [9], and ba- sically consists of full disruption of the nerve fibers con- necting the the afferent signals from the baroreceptors to the brainstem, leading to hypertension, tachycardia, and an increase in blood pressure lability. Twenty days after the surgery, only the increase of blood pressure lability is usually observed [10]. This adaptation had been pre- viously investigated in the light of detrended fluctuation analysis [11]. Blood pressure was recorded from the left femoral artery for 90 minutes in conscious rats. Before the ana- log to digital conversion, blood pressure was low-pass filtered (fc= 50 Hz) for high-frequency noise removal, and recorded with a 2kHz sampling frequency. Diastolic (minimum) values were detected after parabolic interpo- lation and signal artifacts were visually identified and removed. A more detailed account of this experiment can be found in [10]. Since the measurements were done in awake conscious unrestrained rats some distortions in the blood pressure signal arise due to their movements. To reduce this problem we discard series that show any kind of discontinuities. y =x0.9 y =x0.5 101 102 n 103 c) b) ) n ( F 104 103 102 101 100 10-1 105 100 a) ) g H m m ( P B A 95 90 85 80 75 70 65 60 55 0 2000 4000 6000 Counts 8000 10000 10-4 10-3 10-2 10-1 Normalized frequency FIG. 1. Representative data of diastolic blood pressure from one animal. (a) a typical diastolic blood pressure record is presented. (b) The histogram for the blood pressure values. A Gaussian distribution (black dash-line) with sample mean and sample variance was plotted with the histogram, high- lighting the non-Gaussian behavior. (c) DFA for diastolic blood pressure series. For long time scales the DFA expo- nent is α = 0.93. We also applied DFA to shuffled data (red up triangles) obtaining α ≈ 0.5. The curves α = 0.5 (red full line), 0.9 (blue full line) are plotted as guides to the eye. Results and Discussion - Figure 1a a typical diastolic blood pressure record from a control animal is presented. The histogram for the blood pressure values is presented in figure 1c. A Gaussian distribution with sample mean and sample variance was plotted with the histogram, highlighting the non-Gaussian behavior. Furthermore, it clearly presents a positive skew. Despise the fact that the time series has a constant mean of 81.00mmHg over a range of 104 inter-beats, a complex pattern of fluc- tuations were observed. To characterize the stochastic dynamics of the blood pressure a detrended fluctuation analysis (DFA) was performed. The log-log plot of the fluctuation function F (n) is presented in figure 1b. A crossover around n = 35 separate two different behav- iors. For long time scales the DFA exponent α = 0.93 while a for short time scales α = 1.18. This crossover had been discussed previously and it could be associated with autonomic nervous system control of arterial blood pressure [11, 12]. To confirm the presence of long range correlations, the time series were randomly shuffled and DFA was applied. The shuffled time series present DFA exponent α ≈ 0.5, as typical white noise, show at figure 1b in red, which demonstrates that long range correla- tions arises from the blood pressure control system. External perturbations are continuously disrupting the cardiac system. In such noisy environment the autonomic 2 nervous system must keep the blood pressure at accept- able levels by integrating chemical and mechanical input from afferents to regulate the blood pressure. However, the neural transmission itself is noisy [13]. Information transmission along the axons is made by electrical signals. Those signals, called action potentials, are created by ion channels in a cell membrane. They travel down the axon to its end where the neurotransmitters are release in the synaptic gap. Those transmitters activate receptors in the post synaptic neuron [14]. The noise source are di- verse, for example, they could have physical background in thermodynamic or quantum mechanics, as it happens with sensor neurons, or could be built up in the cellular network [15]. In this sense, the cardiac neural control system is under intrinsic and extrinsic noise. 100 10-1 10-2 10-3 ) . u . a ( y c n e u q e r f d e z i l a m r o N 100 10-1 10-2 10-3 ) . u . a ( y c n e u q e r f d e z i l a m r o N −30 −20 −10 0 ps 10 20 30 −30 −20 −10 0 ps 10 20 30 (a) Diastolic pressure distribution control rats (b) Diastolic pressure distribu- tion denervated rats FIG. 2. Semi log histogram of rescaled experimental dias- tolic pressure data. Figure 2a present the data of four con- trol rats while figure 2b shows the data of four denervated animals. All data where rescaled ps = k0(p − p)/σ, where k0 = 4, p is the blood pressure data, p is the median and σ is the sample standard deviation. Despite the fact that the animals were unrestrained during the measurement, the rescaled data collapse. In each figure the stable solution of the model where plotted. For the control animals, figure 2a, a q-Gaussian distribution with q = 11/9 were found. For the denervated rats,figure 2b, a almost Gaussian distribution were found with q = 1.04. Synaptic signal transmission is found to be modeled as a diffusive process wherein additive and multiplicative noise plays relevant role to describe the neuronal response [16, 17]. Let p = P − p where P is the blood pressure and p is a measure of central tendency, like the mean or the median. Then the neural control of blood pressure dynamics could be modeled as a Brownian particle under a restoring force of the baroreflex: dp dt = f (p) + g(p)ξ(t) + η(t). (1) where f (p) is the restoring force, g(p) is the diffusion co- efficient, ξ(t) is the multiplicative noise and η(t) is the ad- ditive noise. They both have zero mean and show Marko- vian correlations, hξtξt′ i = M δtt′ and hηtηt′i = Rδtt′. Physiologically η(t) could be interpreted as external per- turbations on cardiac system while g(p)ξ(t) is a noise arising from the neural transmission. Harris and Wolpert [18] propose an unifying optimal control theory for information process in motor systems. The theory is based on a single physiological assump- tion: the neural noise increases in variance with the size of the control signal. This assumption is then made by the model presented in this paper. Here the noise am- plitude is signal dependent and it is proportional to the control system. Several complex systems, like financial and turbulence [19, 20], has a linear drift coefficient and a quadratic state-dependent diffusion coefficient. In those system f (p) = − dU(p) and g2(p) ∝ U (p). Analogously, dp the restoring force of the baroreflex could be expressed as above with U (p) = τ 2 g(p)2.Generally the baroreflex is represented by a sigmoid function [21], f (p) = A− C 1+e−Bp where A, B, C are the model parameters. Integrating the τ h−Ap + C ln(1+eBp) i sigmoid one can found g2(p) = 2 where A < C, A > 0, and B > 0 must hold that g2(p) be strictly positive. B Equation (1) could be rewritten as Fokker-Planck equation using Ito calculus: ∂F (p) ∂t = − ∂[f (p)F (p)] ∂p + ∂[G′(p)F (p)] ∂p + G(p) ∂2[F (p)] ∂p2 (2) where G(p) = R+M g2(p), G′(p) = ∂G(p) and F (p) is the arterial blood pressure distribution. The physiologically viable solution for equation 2 must have F (p → ∞) = 0 as boundary condition. In those conditions, equation 2 has a q-Gaussian distribution as stationary solution [22]: ∂p F (p) = N (cid:2)1 − (1 − q)βg(p)2(cid:3) 1 1−q (3) By replacing equation 3 in equation 2, we obtain: gg ′{(τ +M )[2+βτ (q−1)g2 +2]−βτ (R+M g2)} = 0. (4) B ln( C R If gg ′ = 0 implies f (p) = 0 for some p∗. As A < C and B > 0 then p∗ = − 1 A − 1) > 0. The positive value of p∗ is associated with the continuous action of the sympa- thetic activity when not inhibited. In other words, when the blood pressure is near the central value there is an continuous activity, called sympathetic tone, of the ner- vous system to increase the blood pressure. Otherwise, if gg ′ 6= 0, equation 4 must hold for every p, which implies q = 2 − 1 . If coupling constant τ is larger than multiplicative noise amplitude such as M/τ → 0 then q → 1 and F (p) converge to Gaussian distribution. The multiplicative noise becomes too small and the long tail vanish. On the other hand, if M/τ → ∞ q → 2 and F (p) converges to the Lorentz distribution [23] 1+M/τ and β = 2(1+M/τ ) 3 destroying homeostasis. In spite of q-Gaussian distribu- tion are defined for q < 3, to describe the non-Gaussian behavior observed in figure 1c and keep the physiological feasibility, q must stick in between 1 and 2. To compare the proposed model with the experimen- tally observed diastolic blood pressure data, a rescale transformation where performed: ps = k0(p−p)/σ, where k0 = 4, p is the blood pressure, p is it median and σ is it standard deviation. Once the animals were unre- strained, each BP series could shown different central values and variability. However, when the rescale where performed the data collapse as figure 2 shows. In the same figure the stable solution of the model, equation 3, where plotted. For control animals, figure 2a, the following parameters were used: τ = 21/5, M = 6/5, A = 7/5, B = 2/5, and C = 9/5. These values imply- ing q = 11/9 ≈ 1.22. A value of q = 1.26 ± 0.1 where observed in heart rate variability [24]. Several other non- equilibrium systems presents q = 1.22,for example, fina- cial markets [25], hadron-hadron collisions [26], and geo- logical faults [27]. For surgically disrupted animals, pre- sented at figure 2b, the following parameters were used: τ = 26/5, M = 1/5, A = 1, B = 1/5, and C = 9/5. These values implying q = 1.04, very close to the Gaus- sian distribution (q = 1.0), showing that the non Gauss- sian fluctuations are intrinsicly associated with the con- trol feedback loop at short time scale. To characterize quantitatively the dynamics, a Monte Carlo simulation of the equation 1 were performed for both group. The same parameters of the analitical curves in figure 2 were used. To understand how the model could reproduce the underlying dynamics a DFA of the simu- lated series were compared with the blood pressure real data. As recently discussed DFA could present a biased estimator for the Hurst exponent [28]. Nevertheless it still holding as good methodology to work on real data [29], and any bias that could be introduced in the fluctu- ation function will happens on both time series, the real and the simulated one, being no hindrance to the analy- sis. The results of the DFA for synthetic series and real data is presented in figure 3. The time scales of the blood pressure control feedback loop is commonly analyzed in three different ranges: high , low and very low frequency . The baroreflex plays a significant role at high and low frequency control [30]. After 20 days, since major damages were inflicted on the neural circuit responsible for the baroreflex, the con- trol were achieved by some other redundant mechanism but with larger response delay [11]. Figure 3 shows the model captures the fluctuations behavior in the barore- flex control range for both groups. Time scales larger than n > 500 are in the very low frequency range and the model developed here plays no role. However, except for long time scales, the fluctuation functions reveals a very similar pattern between synthetic and experimen- tal time series. Based on this analysis, we conclude the 1e+06 100000 10000 1000 100 10 1 ) n ( F 0.1 1 10 100 n 1000 10000 (a) DFA control rats 1e+07 1e+06 100000 10000 ) n ( F 1000 100 10 1 0.1 1 10 100 n 1000 10000 (b) DFA denervated rats FIG. 3. Comparison between DFA analysis of the experimen- tal data and DFA of the time series generated by equation 1. The experimental data for four different animals of each group are plotted with open symbols (purple (cid:3), green ◦, black △ and blue ▽). The DFA from simulated series are plotted with red ×. The curves were translated for better visibility, and values of F (n) are vertically shifted. In figure 3a the control group is presented while in 3b the denervated group is presented. The model capture the fluctuations behavior until the n ≈ 500 for both groups. The error bars represent the standard deviation over 1000 simulations. model presented here captures the complex behavior of blood pressure control. A langevin model based on a sigmoidal restoring force with multiplicative noise for the diastolic blood pressure time series was discussed. The stationary solution was a q-Gaussian distribution with q = 11/9 that describes remarkably well the blood pressure time series recorded from femoral artery for 30 minutes in conscious uncon- strained rats. To investigate the model dynamics a syn- thetic time series was generated by a monte carlo simula- tion of the equivalent Langevin equation. The synthetic time series and the experimental data was characterized using DFA and the fluctuation functions were compared. The fluctuation functions reveals a very similar pattern in the synthetic and experimental time series for high and low frequency. [1] L. K. McCorry, Am. J. Pharm. Educ. 71, 78 (2007). 4 [2] A. C. Guyton Hall, Textbook of Medical Physiology (W.B. Saunders Com- pany, 2000). and J. E. [3] K. M. Gallagher, P. J. Fadel, S. A. Smith, M. Strømstad, K. Ide, N. H. Secher, and P. B. Raven, Exp. Physiol. 91, 79 (2006). [4] M. E. Crandall and C. M. Heesch, Am. J. Physiol. Regul. Integr. Comp. Physiol. 258, R1417 (1990). [5] G. Grassi, B. M. Cattaneo, G. Seravalle, A. Lanfranchi, and G. Mancia, Hypertension 31, 68 (1998). [6] C. K. Peng, J. Mietus, J. M. Hausdorff, S. Havlin, H. E. Stanley, and A. L. Goldberger, Phys. Rev. Lett. 70, 1343 (1993). [7] K. Kiyono, Z. R. Struzik, N. Aoyagi, F. Togo, and Y. Ya- mamoto, Phys. Rev. Lett. 95, 058101 (2005). [8] C. Tsallis et al., Introduction to nonextensive statistical mechanics (Springer, 2009). [9] E. M. Krieger, Circ. Res. 15, 511 (1964). [10] P. P. d. S. Soares, C. S. Port, F. M. F. Abdalla, R. N. De La Fuente, E. D. Moreira, E. M. Krieger, and M. C. Irigoyen, J. Cardiovasc. Pharmacol. 47, 331 (2006). [11] C. E. C. Galhardo, T. J. P. Penna, M. A. de Menezes, and P. Soares, New J. Phys. 11, 103005 (2009). [12] K. Fuchs, A. Y. Schumann, A. Kuhnhold, P. Guzik, J. Piskorski, G. Schmidt, and J. W. Kantelhardt, in Proc. 6th Conference of the European Study Group on Cardiovascular Oscillations, Vol. 56 (2010). [13] R. B. Stein, E. R. Gossen, and K. E. Jones, Nat. Rev. Neurosci. 6, 389 (2005). [14] J. A. White, J. T. Rubinstein, and A. R. Kay, Trends Neurosci. 23, 131 (2000). [15] A. A. Faisal, L. P. Selen, and D. M. Wolpert, Nat. Rev. Neurosci. 9, 292 (2008). [16] C. Boucsein, T. Tetzlaff, R. Meier, A. Aertsen, and B. Naundorf, J. Neurosci. 29, 1006 (2009). [17] M. J. Richardson and R. Swarbrick, Phys. Rev. Lett. 105, 178102 (2010). [18] C. M. Harris and D. M. Wolpert, Nature 394, 780 (1998). [19] J. Peinke, R. Friedrich, et al., Phys. Rev. Lett. 83, 5495 (1999). [20] S. Ghashghaie, W. Breymann, J. Peinke, P. Talkner, and Y. Dodge, Nature 381, 767 (1996). [21] B. Kent, J. Drane, B. Blumenstein, and J. Manning, Cardiology 57, 295 (1972). [22] L. Borland, Physics Letters A 245, 67 (1998). [23] S. Umarov, C. Tsallis, and S. Steinberg, Milan J. Math. 76, 307 (2008). [24] G. Pavlos, L. Karakatsanis, M. Xenakis, E. Pavlos, A. Il- iopoulos, and D. Sarafopoulos, Physica A 395, 58 (2014). [25] M. Ausloos and K. Ivanova, Phys. Rev. E 68, 046122 (2003). [26] G. Barnafoldi, K. Urmossy, and T. Bir´o, in Journal of Physics: Conference Series, Vol. 270 (IOP Publishing, 2011) p. 012008. [27] F. Vallianatos, Europhys. Lett. 102, 28006 (2013). [28] R. Bryce and K. Sprague, Sci. Rep. 2 (2012). [29] Y.-H. Shao, G.-F. Gu, Z.-Q. Jiang, W.-X. Zhou, and D. Sornette, Sci. Rep. 2 (2012). [30] C. Cerutti, C. Barres, and C. Paultre, Am. J. Physiol. Heart Circ. Physiol. 266, H1993 (1994).
1607.01959
1
1607
2016-07-07T10:54:53
The Energy Landscape of Neurophysiological Activity Implicit in Brain Network Structure
[ "q-bio.NC" ]
A critical mystery in neuroscience lies in determining how anatomical structure impacts the complex functional dynamics of human thought. How does large-scale brain circuitry constrain states of neuronal activity and transitions between those states? We address these questions using a maximum entropy model of brain dynamics informed by white matter tractography. We demonstrate that the most probable brain states -- characterized by minimal energy -- display common activation profiles across brain areas: local spatially-contiguous sets of brain regions reminiscent of cognitive systems are co-activated frequently. The predicted activation rate of these systems is highly correlated with the observed activation rate measured in a separate resting state fMRI data set, validating the utility of the maximum entropy model in describing neurophysiologial dynamics. This approach also offers a formal notion of the energy of activity within a system, and the energy of activity shared between systems. We observe that within- and between-system energies cleanly separate cognitive systems into distinct categories, optimized for differential contributions to integrated v.s. segregated function. These results support the notion that energetic and structural constraints circumscribe brain dynamics, offering novel insights into the roles that cognitive systems play in driving whole-brain activation patterns.
q-bio.NC
q-bio
The Energy Landscape of Neurophysiological Activity Implicit in Brain Network Structure Shi Gu1,2, Matthew Cieslak3, Benjamin Baird4, Sarah F. Muldoon5, Scott T. Grafton3, Fabio Pasqualetti6, Danielle S. Bassett2,7,* 1 Applied Mathematics and Computational Science, University of Pennsylvania, Philadelphia, PA, 19104 USA 2 Department of Bioengineering, University of Pennsylvania, Philadelphia, PA, 19104 USA 3 Department of Psychological and Brain Sciences, University of California, Santa Barbara, CA, 93106 USA 4 Center for Sleep and Consciousness, University of Wisconsin - Madison, Madison, WI 53706 5 Department of Mathematics and CDSE Program, University at Buffalo, SUNY, Buffalo, NY 14260 6 Department of Mechanical Engineering, University of California, Riverside, CA, 92521 USA 7 Department of Electrical & Systems Engineering, University of Pennsylvania, Philadelphia, PA, 19104 USA * To whom correspondence should be addressed: [email protected]. Abstract A critical mystery in neuroscience lies in determining how anatomical structure impacts the complex functional dynamics of human thought. How does large-scale brain circuitry constrain states of neuronal activity and transitions between those states? We address these questions using a maximum entropy model of brain dynamics informed by white matter tractography. We demonstrate that the most probable brain states -- characterized by minimal energy -- display common activation profiles across brain areas: local spatially-contiguous sets of brain regions reminiscent of cognitive systems are co-activated frequently. The predicted activation rate of these systems is highly correlated with the observed activation rate measured in a separate resting state fMRI data set, validating the utility of the maximum entropy model in describing neurophysiologial dynamics. This approach also offers a formal notion of the energy of activity within a system, and the energy of activity shared between systems. We observe that within- and between-system energies cleanly separate cognitive systems into distinct categories, optimized for differential contributions to integrated versus segregated function. These results support the notion that energetic and structural constraints circumscribe brain dynamics, offering novel insights into the roles that cognitive systems play in driving whole-brain activation patterns. Author Summary How does the complex interconnection structure between large-scale brain regions impact how we think? Does this structure inform how we move between different PLOS 1/21 thoughts or actions? We address these questions using a simple mathematical model of brain dynamics that is informed by empirical estimates of anatomical interconnection structure between brain regions. Our results suggest that while interconnection structure plays an important role in constraining and predicting brain dynamics, an additional layer of explanatory power is offered by considering energetic constraints on those dynamics. These data offer a richer landscape of mechanisms that enhance our understanding of how we may move from one thought to the next. Introduction A human's adaptability to rapidly changing environments depends critically on the brain's ability to carefully control the time within (and transitions among) different states. Here, we use the term state to refer to a pattern of activity across neurons or brain regions [1]. The recent era of brain mapping has beautifully demonstrated that the pattern of activity across the brain or portions thereof [2] differs in different cognitive states [3]. These variable patterns of activity have enabled the study of cognitive function via the manipulation of distinct task elements [3], the combination of task elements [4, 5], or the temporal interleaving of task elements [6, 7]. Such methods for studying cognitive function are built on the traditional view of mental chronectomy [8], which suggests that brain states are additive and therefore separable in both space and time (although see [9] for a discussion of potential caveats). Philosophically, the supposed separability and additivity of brain states suggests the presence of strong constraints on the patterns of activations that can be elicited by the human's environment. The two most common types of constraints studied in the literature are energetic constraints and structural constraints [10]. Energetic constraints refer to fundamental limits on the evolution [11] or usage of neural systems [12], which inform the costs of establishing and maintaining functional connections between anatomically distributed neurons [13]. Such constraints can be collectively studied within the broad theory of brain function posited by the free energy principal -- a notion drawn from statistical physics and information theory -- which states that the brain changes its state to minimize the free energy in neural activity [14, 15]. The posited preference for low energy states motivates an examination of the time within and transitions among local minimums of a predicted energy landscape of brain activity [16, 17]. While energetic costs likely form critical constraints on functional brain dynamics, an arguably equally important influence is the underlying structure and anatomy linking brain areas. Intuitively, quickly changing the activity of two brain regions that are not directly connected to one another by a structural pathway may be more challenging than changing the activity of two regions that are directly connected [13, 18]. Indeed, the role of structural connectivity in constraining and shaping brain dynamics has been the topic of intense neuroscientific inquiry in recent years [19 -- 23]. Evidence suggests that the pattern of connections between brain regions directly informs not only the ease with which the brain may control state transitions [24], but also the ease with which one can externally elicit a state transition using non-invasive neurostimulation [25]. While energy and anatomy both form critical constraints on brain dynamics, they have largely been studied in isolation, hampering an understanding of their collective influence. Here, we propose a novel framework that combines energetic and structural constraints on brain state dynamics in a free energy model explicitly informed by structural connectivity. Using this framework, we map out the predicted energy landscape of brain states, identify local minima in the energy landscape, and study the profile of activation patterns present in these minima. Our approach offers fundamental insights into the distinct role that brain regions and larger cognitive systems play in PLOS 2/21 distributing energy to enable cognitive function. Further, the results lay important groundwork for the study of energy landscapes in psychiatric disease and neurological disorders, where brain state transitions are known to be critically altered but mechanisms driving these alterations remain far from understood [26, 27]. Fig 1. Conceptual Schematic. (A) A weighted structural brain network represents the number of white matter streamlines connecting brain regions. (B) We model each brain region as a binary object, being either active or inactive. (C) Using a maximum entropy model, we infer the full landscape of predicted activity patterns -- binary vectors indicating the regions that are active and the regions that are not active -- as well as the energy of each pattern (or state). We use this mathematical framework to identify and study local minima in the energy landscape: states predicted to form the foundational repertoire of brain function. Materials and Methods Human DSI Data Acquisition and Preprocessing Diffusion spectrum images (DSI) were acquired for a total of 48 subjects (mean age 22.6±5.1 years, 24 female, 2 left handed) along with a T 1 weighted anatomical scan at each scanning session [28]. Of these subjects, 41 were scanned ones, 1 was scanned twice, and 6 were scanned three times, for a total of 61 scans. DSI scans sampled 257 directions using a Q5 half shell acquisition scheme with a maximum b value of 5000 and an isotropic voxel size of 2.4mm. We utilized an axial acquisition with the following parameters: T R = 11.4s, T E = 138ms, 51 slices, FoV (231,231,123 mm). All participants volunteered with informed written consent in accordance with the Institutional Review Board/Human Subjects Committee, University of California, Santa Barbara. DSI data were reconstructed in DSI Studio (www.dsi-studio.labsolver.org) using PLOS 3/21 State 1State 2State 3Brain NetworkAnteriorPosteriorInferiorSuperior1: Active0: Not activeEdge: Structural ConnectionActivityState EnergyACBActivation Magnitude (a.u.)01 q-space diffeomorphic reconstruction (QSDR) [29]. QSDR first reconstructs diffusion weighted images in native space and computes the quantitative anisotropy (QA) in each voxel. These QA values are used to warp the brain to a template QA volume in MNI space using the SPM nonlinear registration algorithm. Once in MNI space, spin density functions were again reconstructed with a mean diffusion distance of 1.25 mm using three fiber orientations per voxel. Fiber tracking was performed in DSI Studio with an angular cutoff of 55◦, step size of 1.0 mm, minimum length of 10 mm, spin density function smoothing of 0.0, maximum length of 400 mm and a QA threshold determined by DWI signal in the CSF. Deterministic fiber tracking using a modified FACT algorithm was performed until 100, 000 streamlines were reconstructed for each individual. Structural Network Construction Anatomical scans were segmented using FreeSurfer [30] and parcellated according to the Lausanne 2008 atlas included in the connectome mapping toolkit [31]. A parcellation scheme including 234 regions was registered to the B0 volume from each subject's DSI data. The B0 to MNI voxel mapping produced via QSDR was used to map region labels from native space to MNI coordinates. To extend region labels through the gray/white matter interface, the atlas was dilated by 4mm. Dilation was accomplished by filling non-labeled voxels with the statistical mode of their neighbors' labels. In the event of a tie, one of the modes was arbitrarily selected. Each streamline was labeled according to its terminal region pair. From these data, we built structural brain networks from each of the 61 diffusion spectrum imaging scans. Consistent with previous work [13, 24, 25, 32 -- 37], we defined these structural brain networks from the streamlines linking N = 234 large-scale cortical and subcortical regions extracted from the Lausanne atlas [31]. We summarize these estimates in a weighted adjacency matrix A whose entries Aij reflect the structural connectivity between region i and region j (Fig. 1A). Following [24], here we use an edge weight definition based on the quantitative anisotropy (QA). QA is described by Yeh et. al (2010) as a measurement of the signal strength for a specific fiber population a in an ODF Ψ(a) [38, 39]. QA is given by the difference between Ψ(a) and the isotropic component of the spin density function (SDF, ψ) ISO (ψ) scaled by the SDF's scaling constant. Along-streamline QA was calculated based on the angles actually used when tracking each streamline. Although along-streamline QA is more specific to the anatomical structure being tracked, QA is more sensitive to MRI artifacts such as B1 inhomogeneity. QA is calculated for each streamline. We then averaged values over all streamlines connecting a pair of regions, and used this value to weight the edge between the regions. Resting state fMRI data As an interesting comparison to the computational model, we used resting state fMRI data collected from an independent cohort composed of 25 healthy right-handed adult subjects (15 female), with a mean age of 19.6 years (STD 2.06 year). All subjects gave informed consent in writing, in accordance with the Institutional Review Board of the University of California, Santa Barbara. Resting-state fMRI scans were collected on a 3.0-T Siemens Tim Trio scanner equipped with high performance gradients at the University of California, Santa Barbara Brain Imaging Center. A T2*-weighted echo-planar imaging (EPI) sequence was used (TR=2000 ms; TE=30 ms; flip angle=90°; acquisition matrix=64×64; FOV=192 mm; acquisition voxel size = 3×3×3.5 mm; 37 interleaved slices; acquisition length=410s). PLOS 4/21 We preprocessed the resting state fMRI data using an in-house script adapted from the workflows described in detail elsewhere [40, 41]. The first four volumes of each sequence were dropped to control for instability effects of the scanner. Slice timing and motion correction were performed in AFNIusing 3dvolreg and FreeSurfer's BBRegister was used to co-register functional and anatomical spaces. Brain, CSF, and WM masks were extracted, the time series were masked with the brain mask, and grand-mean scaling was applied. The temporal derivative of the original 6 displacement and rotation motion parameters was obtained and the quadratic term was calculated for each of these 12 motion parameters, resulting in a total of 24 motion parameters which were regressed from the signal. The principal components of physiological noise were estimated using CompCor (aCompCor and tCompCor) and these components were additionally regressed from the signal. The global signal was not regressed. Finally, signals were low passed below 0.1 Hz and high passed above 0.01 Hz in AFNI. To extract regional brain signals from the voxel-level time series, a mask for each brain region in the Lausanne2008 atlas was obtained and FreeSurfer was used to individually map regions to the subject space. A winner-takes-all algorithm was used to combine mapped regions into a single mask. The resulting signal for each region was then extracted in FreeSurfer using mrisegstats. Following data preprocessing and time series extraction, we next turned to extracting observed brain states. Importantly, physiological changes relating to neural computations take place on a time scale much smaller than the time scale of BOLD image acquisition. Thus, we treat each TR as representing a distinct brain state. To maximize consistency between the model-based and data-based approaches, we transformed the continuous BOLD magnitude values into a binary state vector by thresholding regional BOLD signals at 0. From the set of binary state vectors across all TRs, we defined activation rates in a manner identical to that described for the maximum entropy model data. Defining an Energy Landscape We begin by defining a brain state both intuitively and in mathematical terms. A brain state is a macroscopic pattern of BOLD activity across K regions of the brain (Fig. 1A). For simplicity, here we study the case in which each brain region i can be either active (σi = 1) or inactive (σi = 0). Then, the binary vector σ = (σ1, σ2, . . . , σK) respresents a brain state configuration. Next, we wish to define the energy of a brain state. We build on prior work demonstrating the neurophysiological relevance of maximum entropy models in estimating the energy of brain states in rest and task conditions [42, 43]. For a given state σ, we write its energy in the second order expansion: Jiσi, (1) (cid:88) i(cid:54)=j Jijσiσj −(cid:88) i E(σ) = − 1 2 where J represents an interaction matrix whose elements Jij indicate the strength of the interaction between region i and region j. If Jij > 0, this edge (i, j) decreases the energy of state σ, while if Jij < 0, this edge (i, j) increases the energy of state σ. The column sum of the structural brain network, Ji =(cid:80) K, is the strength of √ j Jij/ region i. In the thermodynamic equilibrium of the system associated with the energy defined in Eqn [1], the entropy of the system is maximized and the probability of the configuration σ is P (σ) ∝ e−E(σ). The choice of the interaction matrix is an important one, particularly as it tunes the relative contribution of edges to the system energy. In this study, we seek to study structural interactions in light of an appropriate null model. We therefore define the PLOS 5/21 interaction matrix to be equal to the modularity matrix [44] of the structural brain network: for i (cid:54)= j, where A is the adjacency matrix, pi =(cid:80)K (A − ppT/2m)ij i=1 Aij, and 2m =(cid:80)K 1 2m Jij = choice ensures that any element Jij measures the difference between the strength of the edge Aij in the brain and the expected strength of that edge in an appropriate null model (here given as the Newman-Girvan null model [45]). If the edge is stronger than expected, it will decrease the energy of the whole system when activated, while if the edge is weaker than expected, it will increase the energy of the whole system when activated. (2) j=1 pj. This Discovering Local Minima The model described above provides an explicit correspondence between a brain's state and the energy of that state, in essence formalizing a multidimensional landscape on which brain dynamics may occur. We now turn to identifying and characterizing the local minima of that energy landscape (Fig. 1C). We begin by defining a local minimum: K) is a local minimum if E(σ) ≥ E(σ∗) for all vectors σ a binary state σ∗ = (σ∗ satisfying σ − σ∗1 = 1, which means that the state σ∗ realizes the lowest energy among its neighboring states within the closed unit sphere. We wish to collect all local minima in a matrix Σ∗ with 1, . . . , σ∗  Σ∗ =  1,2 2,2 1,1 2,1 σ∗ σ∗ σ∗ σ∗ ... ... K,1 σ∗ σ∗ K,2 ··· ··· . . . ··· 1,N 2,N σ∗ σ∗ ... σ∗ K,N (3) K×N where N is the number of local minima and K is the number of nodes in the structural brain network (or equivalently the cardinality of the adjacency matrix A). Now that we have defined a local minimum of the energy landscape, we wish to discover these local minima given the pattern of white matter connections represented in structural brain networks. To discover local minima of E(σ), we first note that the total number of states σ = (σ1, . . . , σK) is 2K, which -- when K = 234 -- prohibits an exhaustive analysis of all possibilities. Moreover, the problem of finding the ground state is an NP-complete problem [46], and thus it is unrealistic to expect to identify all local minima of a structural brain network. We therefore choose to employ a clever heuristic to identify local minima. Specifically, we combine the Metropolis Sampling method [47] and a steep search algorithm using gradient descent methods. We identify a starting position by choosing a state uniformly at random from the set of all possible states. Then, we step through the energy landscape via a random walk driven by the Metropolis Sampling criteria (see Algorithm 1). At each point on the walk, we use the PLOS 6/21 steep search algorithm to identify the closest local minimum. Algorithm 1: Heuristic Algorithm to Sample the Energy Landscape to Identify Local Minima. 1 Let σj be the vector obtained by changing the value of the j-th entry of σ; 2 for t = 1 : N do 3 t−1 with probability p = min(1, e−β∗(E( σ)−E(σt−1))) and Randomly select an index j ∈ {1, ..., K}; Set σt = σt = σj σt = σt = σt−1 otherwise; while σt is not a local minimum do Set j∗ = arg minj E( σj t ); Set σt = σj∗ t ; 4 5 6 7 8 end Set σ∗ t = σt; 9 10 end 1, σ∗ 2, . . . , σ∗ Here, σ1, σ2, . . . , σN are the sampled states, σ∗ N are the sampled local minima, and β is the temperature parameter which can be absorbed in E(σ). In the context of any sampling procedure, it is important to determine the number of samples necessary to adequately cover the space. Theoretically, we wish to identify a number of samples N following which the distribution of energies of the local minima remains stable. Practically speaking, we choose 4 million samples in this study, and demonstrate the stability of the energy distribution in the Supplement. A second important consideration is to determine the initial state, that is the state from which the random walk begins. Here we choose this state uniformly at random from the set of all possible states. However, this dependence on a uniform probability distribution may not be consistent with the actual probability of states in the energy landscape. We therefore must ensure that our results are not dependent on our choice of initial state. To ensure independence from the initial state, we dismiss the first 30, 000 local minima identified, and we demonstrate in the Supplement that this procedure ensures our results are not dependent on the choice of initial state. Characterizing Local Minima Following collection of local minima, we wished to characterize their nature as well as their relationships to one another. First, we estimated the radius of each local minimum as the Hamming distance from the minimum to the closest sampled point on energy landscape. Next, calculated the Hamming distance from each local minimum to the first sampled local minimum, a second quantification of the diversity of the energy lanscape that we traverse in our sampling. Finally, we quantify how diverse the observed local minima are by calculating the pairwise normalized mutual information [48] of each pair of local minima. Next, we wished to understand the role of different regions and cognitive systems in the minimal energy states. Cognitive systems are sets of brain regions that show coordinated activity profiles in resting state or task-based BOLD fMRI data [9]. They include the visual, somatosensory, motor, auditory, default mode, salience, fronto-parietal, cingulo-opercular, dorsal and ventral attention systems, as well as subcortical areas. Here, the specific association of regions of interest to cognitive systems are exactly as listed in [24] and based originally on results described in [49]. We characterize the roles of these systems in the local minima by assessing their activation rates, as well as the utilization energies required for communication within and between systems. PLOS 7/21 Activation Rates Intuitively, we define the activation rate of a node i as the average activation state of that node over all the local minima. Formally, the activation rate for region i is defined as ri = , (4) (cid:80)N l=1 σ∗ N il where l indexes over states, and recall N is the number of local minima. The computed activation rate offers a prediction of which regions are more versus less active across the local minima (that is, the brain's locally "stable " states), and can be directly compared with the resting state activation rate estimated from empirical fMRI data. Utilization Energies To complement the information provided by the activation rates, we also defined the energetic costs associated with utilizing within- and between-systems interactions. We note that each cognitive system is a subnetwork of the whole brain network. We use the index set I to indicate the set of nodes associated with the cognitive system, and thus I gives the number of nodes in the system. Then, for a given state σ, the within-system energy measures the cost associated with the set of interactions constituting the subnetwork. The between-system energy measures the cost associated with the set of interactions between the subnetwork and all other nodes in the whole network. Formally, we define EW (σ) = − 1 2I(I − 1) EB(σ) = − 1 2I(K − I)  (cid:88)  (cid:88) i(cid:54)=j,i,j∈I i∈I,j /∈I   Jijσiσj Jijσiσj where EW measures the within-system energy, EB measures the between-system energy, and the normalization coefficients 1/(II − 1), 1/(I(K − I)) are chosen by considering the number of the corresponding interactions. Permutation Tests for State Association For a given local minimum configuration σ∗, we associate it with system iσ∗ , iσ∗ = arg max i NMI(σ∗, σsys i ), i where σsys is the configuration pattern of system i such that the corresponding regions for system i are activated and the others not and where "NMI" refers to the Normalized Mutual Information [48], which is used to measure the similarities between the two states. To obtain the null distribution, for each local minimum configuration σ∗ in the collection Σ∗, we permute the configuration at each position of σ∗ to achieve a random configuration with the same activation rate, and we then compute the associated percentage in each system. Then we repeat this procedure to generate N samples and construct the null distribution of the probability of being configured as each system pattern. Considering the large size of the state collection, the variance of the samples in the null distribution will be small. We pick N = 50 here. See Fig. 4 D for the results. PLOS 8/21 Results Local Minima in the Brain's Energy Landscape By sampling the energy landscape of each structural connectivity matrix, we identified an average of approximately 450 local minima or low energy brain states: binary vectors indicating the pattern of active and inactive brain regions (see Methods). On average across the 61 scans, 144 brain regions were active in a given local minimum, representing 61.70% of the total (standard deviation: 6.04%). This large percentage suggests that the brain may utilize a broad set of rich and diverse activations to perform cognitive functions [50], rather than activation patterns localized to small geographic areas. To quantify this diversity, we examined the location of minima on the energy landscape, the size of the basins surrounding the minima, and the mutual information between minima. First, we estimated the distance from the first local minima identified to all subsequent minima (see Methods; Fig. 2A). We observe an order of magnitude change in the distance between the first and second local minima, and the first and last local minima, suggesting that local minima span a broad geographic domain in the energy landscape. Interestingly, these minima differ not only in their location on the energy landscape, but also in the size of the basins surrounding them. We estimate basin size by calculating the radius of each local minimum (see Methods) and show that the distribution of radii follows a power-law, with the majority of minima displaying a small radius, and only a few minima displaying a large radius (Fig. 2B). More specifically, we fit the function P (r) = Cr−α -- where C is a constant -- to the data using a statistically principled approach [51, 52]. We identified an α = 2.6300 for r > 6, and the p-value for the goodness of fit was p < 1 × 10−5 indicating that the power law was an accurate fit to the data. As a final quantification of minima diversity, we estimated the normalized mutual information between every pair of local minima, as an intuitive measurement of the similarity between anatomical compositions of the minima. We observe that the probability distribution of normalized mutual information between minima pairs is heavy tailed, indicating that most minima pairs display very dissimilar anatomical compositions, and only a few minima pairs display similar anatomical compositions (Fig. 2C). From a neurophysiological perspective, it is also important to note that these local minima displayed significant local structure. Specifically, we found that regions within known cognitive systems tended to be active together. The probability that regions were co-active is 48.22%, which was significantly greater than that expected in a null distribution (associated p − value was p < 1 × 10−5; see Methods). These results indicate that the structural connectivity between brain regions, and the assumption of energy conservation, predict that regions that belong to the same cognitive system will tend to be co active with one another during diverse cognitive functions. Indeed, these predictions are consistent with previous studies of functional neuroimaging data demonstrating that groups co-active regions tended to align well with known cognitive systems [53, 54]. Activation Rates of Cognitive Systems Given the alignment of activation patterns with cognitive systems, we next asked whether certain cognitive systems were activated more frequently than others. To address this question, we studied the activation rate of each cognitive system, which measures how frequently the regions in the cognitive system participated in the set of states identified as local minima. Intuitively, if the activation rate is high, the system is more likely to be active in diverse brain states. We observed that systems indeed showed signficantly different activation rates (Fig. 2D). Sensorimotor systems (auditory, PLOS 9/21 Fig 2. Simulated Activation Rates.(A) The distribution of distances from the first local minimum to other local minima. Each point and error-bar is calculated across a bin of 30 minima; error bars indicate standard error of the mean over the 30 minima. (B) The probability distribution of the radius of each local minimum is well-fit by a power-law. The radius of a local minimum is defined as its distance to the closest sampled point on the energy landscape. (C) The distribution of the pairwise normalized mutual information between all pairs of local minima. (D) Average activation rates for all 14 a priori defined cognitive systems [49]. Error bars indicate standard error of the mean across subjects. visual, somatosensory) tended to display the lowest activation rates, followed by higher order cognitive systems (salience, attention, fronto-parietal, and cingulo-opercular), and subcortical structures. The system with the largest activate rate was the default mode system, suggesting that activation of this system is particularly explicable from structural connectivity and the assumption of energy conservation. The unique role of the default mode system is consistent with predictions from network control theory that highlight the optimal placement of default mode regions within the network to maximize potential to move the brain into many easily reachable states with minimal energetic costs [24]. It is important to determine whether this activation rate is driven by simple properties of the structural connectivity matrix that do not depend on assumptions of energy conservation. To address this question, we next assessed the relationship between a simple summary statistic of the structural connectivity matrix -- the strength, or weighted degree, of a brain region -- and the predicted activation rate drawn from the maximum entropy model. We observed that the activation rate was not well predicted by the weighted degree on average over brain regions (see Supplement). These data suggest that the additional assumption of energy conservation produces a set of brain states that cannot be predicted from simple statistics of structural connectivity alone. PLOS 10/21 x100101102Pr(X ≥ x)10-510-410-310-210-1100The n-th local Minimum105045165225285345405465525585s-s020406080100120140160180200Normal Mutual Information00.20.40.60.81Probability00.050.10.15ABCActivation Rate0.500.550.600.650.700.750.80Ventral AttentionCingulo-OpercularFronto-ParietalDefault ModeSomatosensoryDorsal AttentionVisualOtherAuditorySubcorticalD Relating Predicted Activation Rates to Rates Observed in Functional Neuroimaging Data Before exercising the model further to explore how energy is utilized in the brain, we wished to quantify the relationships between the theoretically predicted activation rates, and activation rates observed empirically in functional neuroimaging data. Specifically, we studied fMRI data acquired in a separate set of healthy adult human subjects. Next, we wished to directly quantify similarities between the predicted activation rates and those observed empirically in resting state fMRI. In the resting state data, we observed that highly active regions were located in broad swaths of frontal and parietal cortex, as well as medial prefrontal, precuneus, and cingulate (Fig. 3A). This pattern of high activation is consistent with the so-called "default-mode" of resting state brain function [55]. In our maximum entropy model, we observed that the areas predicted to have high activation rates show a broad similarity to those observed empirically in the resting state (Fig. 3B). Indeed, we observed that the empirical resting activation rate of brain regions is significantly correlated with the activation rate predicted from the maximum entropy model (Fig. 3C; Pearson correlation coefficient r = 0.18, p = 0.0046). These results suggest that the modeling framework we use here has significant similarities to observable features of resting state brain dynamics. However, it is important to mention that there are also noticeable differences between the two maps: the predicted activation rates are strong along the medial wall, while the resting state activation rates extend to larger sections of lateral cortices. Fig 3. Validating Predicted Activation Rates in Functional Neuroimaging Data. (A) From resting state BOLD data acquired in an independent cohort, we estimated the true activation rate by transforming the continuous BOLD magnitudes to binary state vectors by thresholding the signals at 0 (see Methods). We use these binary state vectors to estimate the activation rates of each brain region across the full resting state scan. Here we show the mean activation rate of each brain region, averaged over subjects. (B) For comparison, we also show the mean predicted activation rate estimated from the local minima of the maximum entropy model, as defined in Equation [4], and averaged over subjects. (C) We observe that the activation rates estimated from resting state fMRI data are significantly positively correlated with the activation rates estimated from the local minima of the maximum entropy model (Pearson's correlation coefficient r = 0.18, p = 0.0046). Each data point represents a brain region, with either observed or predicted activation rates averaged over subjects. Utilization Energies of Cognitive Systems We next turned to exercising our model to further understand the potential constraints on brain state dynamics. Specifically, we asked how cognitive systems utilized the PLOS 11/21 Model Activation RateSimulated Activation RateReal Activation Rate0.40.50.60.70.80.91.0AReal Activation Rate0.480.4850.490.4950.50.5050.510.5150.52r = 0.1845p = 0.0046BC0101 minimal energy presumably available to them. Intuitively, this question encompasses both how energy is utilized by within-system interactions, and how energy is utilized by between-system interactions. We therefore defined the within-system energy, which measures the cost associated with the set of interactions constituting the cognitive system, and the between-system energy, which measures the cost associated with the set of interactions between cognitive systems. We observed a fairly strong dissociation between these two variables: cognitive systems that display a large within-system energy are not necessarily those that display a large between-system energy (see Fig. 4A and B). Indeed, within- and between-system energies are not significantly correlated across systems (Pearson's correlation coefficient r = 0.2287 and p = 0.5250), suggesting that these two variables offer markers of distinct constraints. Moreover, we observed that the 2-dimensional plane mapped out by the within- and between-system energies of all brain regions revealed the presence of 4 surprisingly distinct clusters (Fig. 4C) that are not explicable by simple statistics such as network degree (see Supplement). Each cluster represents a unique strategy in energy utilization that is directly reflected in its activation pattern; in other words, each cluster offers a distinct balance between the energetic costs of within-system interactions and the energetic costs of between-system interactions. The central cluster, displaying high within-system energies but low between-system energies, is composed of subcortical and fronto-parietal systems. A high between-system energy cone emanating from this central cluster is composed of predominantly primary and secondary sensorimotor cortices in somatosensory, visual, and auditory systems. A second cone emanating from the central cluster with a slightly lower between-system energy is composed predominantly of regions in the default mode system. The final cone emanating from the central cluster with between-system energies near zero is composed predominantly of regions in the dorsal and ventral attention systems. These results suggest that sensorimotor, default mode, attention, and cognitive control circuits display differential preferences for energy utilization: regions in attentional systems share less energy with other networks than regions in sensorimotor systems, while the default mode maintains an intermediate balance. The clear differences in the energies utilized by different cognitive systems and by between-system interactions begs the question of whether the brain cares about these energies. Does the brain prefer smaller within-system energies, smaller between-system energies, or some balance between the two? To address this question, we studied the ensemble of local minima, and asked which systems were commonly expressed. Specifically, for each local minimum, we determined which system was most activated, assigned the minimum to that system, and performed this assignment for all minima. We observed that 3 systems were represented at higher percentages than expected in a permutation-based null model (see Methods): the default mode system, the visual system, and the somatosensory system (Fig. 4D). Importantly, these three systems represent the systems with the highest between system energies (Fig. 4C), but are indistinguishable from other cognitive systems in terms of within-system energy. These results suggest that the brain may prefer high integration between systems over low integration, and that the constraint of between-system energies is more fundamental to brain function than the constraint of within-system energies. Discussion In this paper, we address the question of how large-scale brain circuitry and distinct energetic constraints produce whole-brain patterns of activity. We build our approach on a maximum entropy model of brain dynamics that is explicitly informed by estimates of white matter microstructure derived from deterministic tractography algorithms. The PLOS 12/21 Fig 4. Utilization Energies of Cognitive Systems. (A) Average within-system energy of each cognitive system; error bars indicate standard error of the mean across subjects. (B) Average between-system energy of each cognitive system; error bars indicate standard error of the mean across subjects. (C) The 2-dimensional plane mapped out by the within- and between-system energies of different brain systems. Each data point represents a different brain region, and visual clusters of regions are highlighted with lightly colored sectors. The sector direction is determined by minimizing the squared loss in point density of the local cloud and the width is determined by the orthogonal standard derivation at the center along the sector direction. In this panel, all data points represent values averaged across subjects. (D) The percentages of minima displaying preferential activation of each system; each minima was assigned to the system which whom it shared the largest normalized mutual information. Errorbars indicate the differences between the observed percentages and those of the null distribution with random activation patterns across regions. model allows us to study minimal energy states, which we observe to be composed of co-activity in local spatially-contiguous sets of brain regions reminiscent of cognitive systems. These systems are differentially active, and activity patterns are significantly correlated with the observed activation rate measured in a separate resting state fMRI data set. Finally, we exercise this model to ask how cognitive systems utilize the minimal energy presumably available to them. We find that the energy utilized within and between cognitive systems distinguishes 4 classes of energy utilization dynamics, corresponding to sensorimotor, default mode, attention, and cognitive control functions. These results suggest that diverse cognitive systems are optimized for differential contributions to integrated versus segregated function via distinct patterns of energy utilization. More generally, the results highlight the importance of considering energetic constraints in linking structural connectivity to observed dynamics of neural activity. PLOS 13/21 -0.4-0.200.20.40.60.81.01.2-1.2-1.0-0.8-0.6-0.4-0.20Ventral AttentionCingulo-OpercularFronto-ParietalDefault ModeSomatosensoryDorsal AttentionVisualOtherAuditorySubcorticalA-0.4-0.200.20.40.60.81.01.2Ventral AttentionCingulo-OpercularFronto-ParietalDefault ModeSomatosensoryDorsal AttentionVisualOtherAuditorySubcorticalBx10-5x10-6-2-1.5-1.0-0.500.51-1.5-1.0-0.500.51.01.52.02.5x10-6x10-5CDWithin-System EnergyWithin-System EnergyBetween-System EnergyBetween-System EnergyVentral AttentionCingulo-OpercularFronto-ParietalDefault ModeSomatosensoryDorsal AttentionVisualAuditorySubcorticalOther-0.4-0.3-0.2-0.100.10.2Relative Activation Rate The Role of Activation vs. Connectivity in Understanding Brain Dynamics As the interest in understanding structural brain connectomics has blossomed in the last several years [56, 57], it has not been accompanied by an equally vivid interest in linking subsequent insights to the more traditional notions of brain activation profiles [58]. Indeed, the fields of systems and cognitive neuroscience have instead experienced a pervasive divide between the relatively newer notions of connectome mapping and the relatively traditional yet highly effective notions of brain mapping [59], which have led to powerful insights into neural function in the last quarter century [60]. This divide is at least in part due to the fact that graph theory and network-based methods on which the field of connectomics is based have few tools available to link node properties (activity) with edge properties (connectivity) [61]. While technically explicable, however, the conceptual divide between these fields can only lead to their detriment, and synergistic efforts are necessary to develop a language in which both activity and connectivity can be examined in concert [58]. This study offers one explicit mathematical modeling framework in which to study the relationships between activation profiles across the whole brain and underlying structural connectivity linking brain regions. Complementary approaches include the model-based techniques formalized in the publicly available resource The Virtual Brain [62, 63]. Co-activation Architecture In this study, we observed that brain regions affiliated with known cognitive systems -- including somatosensory, visual, auditory, default mode, dorsal and ventral attention, fronto-parietal, and cingulo-opercular -- also tend to be active together with one another in low energy brain states. Indeed, these theoretical results are consistent with previous studies of functional neuroimaging data demonstrating that groups of co-active regions tended to align well with known cognitive systems [53, 54]. This correspondence is particularly interesting when one considers how these cognitive systems were initially defined: and that is, based on strong and dense functional connectivity [49]. Thus, our results point to a fundamental mapping between activity and connectivity: regions that are active together tend to be functionally connected together. The presence of such a map has been empirically observed in the resting state (though not in task [58]), in both healthy controls and in people with schizophrenia where the map appears to be fundamentally altered in its nature [64 -- 66]. Here we offer a mechanism for this mapping based on a combined consideration of energy- and connectivity-based constraints. Critical Importance of Energy Constraints The quest to understand and predict brain dynamics from the architecture of underlying structural connectivity is certainly not a new one. In fact, there have been concerted efforts over the last decade and more to identify structural predictors of the resting state BOLD signal. Seminal contributions have included the observations of statistically significant correlations between structural connectivity estimated from diffusion imaging data and functional connectivity estimated from fMRI [67], as well as extensions of these correlations that account for long distance paths along white matter tracts [22] and spectral properties of structural matrices [68]. The question of how brain structure constrains a wide range of brain states (beyond simply the resting state) is a very open area of inquiry. Moreover, this question is particularly challenging to address with empirical data because it is difficult to obtain data from humans in more than a handful of task states [69]. For this reason, computational models play a very important role in offering testbeds for the development of theories linking structure to ensembles of PLOS 14/21 brain states, which can in turn offer testable predictions. Our results suggest that an understanding of the relationship between brain structure and function is perhaps ill-constrained when examining connectivity alone. The additional assumption of energy conservation produces a set of brain states that cannot be simply predicted from statistics of structural connectivity, perhaps offering a mechanism for the large amount of unexplained variance in prior predictions [22, 67]. Methodological Considerations Our results are built on the formalism of the maximum entropy model, which is predicated on pairwise statistics [70]. However, emerging evidence suggests that some neurophysiological phenomenon are better studied in the framework of simplicial complexes rather than dyads [71]. For example, integrate and fire neurons exposed to common fluctuating input display strong beyond-pairwise correlations that cannot be captured by maximum entropy models [72]. Similar arguments can also be made for co-activation patterns in BOLD fMRI [53, 54, 73]. It will be interesting in future to determine the role energetic and structural constraints on observed higher-order functional interactions during human cognitive function. A second important consideration is that the maximum entropy model is appropriate for systems at equilibrium. Therefore, the local minima identified may not accurately represent the full class of states expected to be elicited by daily activity. Instead, the local minima identified here are expected to more accurately represent the set of states expected to appear as a person rests in the so-called default mode of brain function, which is thought to lie near a stable equilibrium [21]. Such an interpretation is consistent with our findings that the activation rates predicted by the maximum entropy model are strongly correlated with the activation rates observed in resting state fMRI data. Conclusion The analyses presented in this study produce information regarding an underlying energy landscape through which the brain is predicted to move. The existence of such a landscape motivates the very interesting question of how the brain transitions between states. In sampling this landscape, we have used a simple random walk in an effort to extract a large ensemble of possible brain states, measured by local minima. However, the question of which walks a healthy (or diseased) brain might take through this landscape remains open. Such walks or dynamical trajectories may be determined by energetic inputs to certain regions of the brain [24], either by external stimuli or by neuromodulation [25]. In this context, network control theory may offer explicit predictions regarding the optimal dynamic trajectories that the brain may take through a set of states to move from an initial state to a target state with little energetic resources [24, 74]. In addition to inputs to single regions, changes in a cognitive task -- for example elicited by task-switching paradigms -- may also drive a specific trajectory of brain states. Indeed, it is intuitively plausible that the asymmetric switch costs observed between cognitively effortful and less effortful tasks [75, 76] may be explained by characteristics of the energy landscape defined by structural connections between task-activated brain regions. PLOS 15/21 Supporting Information S1 Fig. Stability of the Energy Distribution with respect to the Number of Samples. We plot the probability distribution of the energy for the first 2000, 4000, 6000 and 8000 samples. We observe that the shapes of the probability distributions are qualitatively consistent. We confirm this qualitative observation with Kolmogorov-Smirnov tests (see Supplemental text). S2 Fig. Activation Rate is Poorly Predicted by Regional Degree and Energy. (A) Scatterplot of weighted regional degree versus activation rate. (B) Scatterplot of regional energy versus activation rate. We observe that the activation rate is not well predicted by either regional energy or weighted degree. S3 Fig. Relationship Between Utilization Energies and Degree. (A) Scatterplot of within-system connectivity versus between-system connectivity, for individual brain regions that are color-coded by cognitive system. (B) The same data presented in panel (A) except only for the default mode, attention, and task control systems, demonstrating the indistinguishability of default mode and attention systems. (C) Scatterplot of the between- and within- system energy. (D) The same data presented in panel (C) except only for the detault mode, attention, and task control systems. We observe that cognitive systems are more clearly separated in the 2-dimensional space of the within- and between-system energies than in the 2-dimensional space of the within- and between-system connectivity. Across all four panels, data points indicate brain regions, and color of data points indicates which cognitive system that region is affiliated with (see legend for map from color to cognitive system). S4 Fig. Simulated Activation Rate is Significantly Correlated with the Rates Observed in Resting State Functional Neuroimaging Data. We observe that the activation rates estimated from resting state fMRI data are significantly positively correlated with the activation rates estimated from the local minima of the maximum entropy model (Pearson's correlation coefficient r = 0.17, p = 0.0094). Each data point represents a brain region, with either observed or predicted activation rates averaged over subjects. S1 Appendix. S1 Appendix. Supplementary Information for "The Energy Landscape of Neurophysiological Activity Implicit in Brain Network Structure". Acknowledgments We thank Ankit Khambhati, Lia Papadopoulus, and Evelyn Tang for helpful comments on an earlier fersion of this manuscript. D.S.B., S.G., and R.F.B. acknowledge support from the John D. and Catherine T. MacArthur Foundation, the Alfred P. Sloan Foundation, the Army Research Laboratory and the Army Research Office through contract numbers W911NF-10-2-0022 and W911NF-14-1-0679,the National Institute of Mental Health (2-R01-DC-009209-11), the National Institute of Child Health and Human Development (1R01HD086888-01), the Office of Naval Research, and the National Science Foundation (BCS-1441502, BCS-1430087, and PHY-1554488). FP acknowledges support from BCS-1430280. B.B. was supported by the National Institutes of Health under Ruth L. Kirschstein National Research Service Award F32NS089348 from the NINDS. PLOS 16/21 References 1. Tang YY, Rothbart MK, Posner MI. Neural correlates of establishing, maintaining, and switching brain states. Trends in cognitive sciences. 2012;16(6):330 -- 337. 2. Mahmoudi A, Takerkart S, Regragui F, Boussaoud D, Brovelli A. Multivoxel pattern analysis for fMRI data: a review. Comput Math Methods Med. 2012;2012:961257. 3. Gazzaniga MS, editor. The cognitive neurosciences. MIT Press; 2013. 4. Szameitat AJ, Schubert T, Muller HJ. How to test for dual-task-specific effects in brain imaging studies: an evaluation of potential analysis methods. Neuroimage. 2011;54(3):1765 -- 1773. 5. Alavash M, Hilgetag CC, Thiel CM, Giessing C. Persistency and flexibility of complex brain networks underlie dual-task interference. Hum Brain Mapp. 2015;36(9):3542 -- 3562. 6. Ruge H, Jamadar S, Zimmermann U, Karayanidis F. The many faces of preparatory control in task switching: reviewing a decade of fMRI research. Hum Brain Mapp. 2013;34(1):12 -- 35. 7. Muhle-Karbe PS, De Baene W, Brass M. Do tasks matter in task switching? Dissociating domain-general from context-specific brain activity. Neuroimage. 2014;99:332 -- 341. 8. Donders FC. On the speed of mental processes. Acta Psychol. 1969;30:412 -- 431. 9. Mattar MG, Cole MW, Thompson-Schill SL, Bassett DS. PLoS Comput Biol. 2015;11(12):e1004533. 10. Bullmore E, Sporns O. The economy of brain network organization. Nature Reviews Neuroscience. 2012;13(5):336 -- 349. 11. Niven JE, Laughlin SB. Energy limitation as a selective pressure on the evolution of sensory systems. J Exp Biol. 2008;211:1792 -- 1804. 12. Attwell D, Laughlin SB. An energy budget for signalling in the grey matter of the brain. J Cereb Blood Flow and Metab. 2001;21:1133 -- 1145. 13. Bassett DS, Greenfield D, Meyer-Lindenberg A, Weinberger DR, Moore SW, Bullmore ET. Efficient physical embedding of topologically complex information processing networks in brains and computer circuits. PLoS Comput Biol. 2010;6(4):e1000748. 14. Friston K, Kilner J, Harrison L. A free energy principle for the brain. Journal of Physiology-Paris. 2006;100(1):70 -- 87. 15. Friston K. The free-energy principle: a unified brain theory? Nature Reviews Neuroscience. 2010;11(2):127 -- 138. 16. Moreno-Bote R, Rinzel J, Rubin N. Noise-induced alternations in an attractor network model of perceptual bistability. Journal of neurophysiology. 2007;98(3):1125 -- 1139. 17. Tsodyks M, Pawelzik K, Markram H. Neural networks with dynamic synapses. Neural computation. 1998;10(4):821 -- 835. PLOS 17/21 18. Achard S, Bullmore E. Efficiency and cost of economical brain functional networks. PLoS computational biology. 2007;3(2):e17. 19. Honey C, Sporns O, Cammoun L, Gigandet X, Thiran JP, Meuli R, et al. Predicting human resting-state functional connectivity from structural connectivity. Proceedings of the National Academy of Sciences. 2009;106(6):2035 -- 2040. 20. Honey CJ, Thivierge JP, Sporns O. Can structure predict function in the human brain? Neuroimage. 2010;52(3):766 -- 776. 21. Deco G, Jirsa VK. Ongoing cortical activity at rest: criticality, multistability, and ghost attractors. J Neurosci. 2012;32(10):3366 -- 3375. 22. Goni J, van den Heuvel MP, Avena-Koenigsberger A, Velez de Mendizabal N, Betzel RF, Griffa A, et al. Resting-brain functional connectivity predicted by analytic measures of network communication. Proc Natl Acad Sci U S A. 2014;111(2):833 -- 838. 23. Becker CO, Pequito S, Pappas GJ, Miller MB, Grafton ST, Bassett DS, et al. Accurately Predicting Functional Connectivity from Diffusion Imaging. arXiv;1512:02602. 24. Gu S, Pasqualetti F, Cieslak M, Telesford QK, Yu AB, Kahn AE, et al. Controllability of structural brain networks. Nat Commun. 2015;6:8414. 25. Muldoon SF, Pasqualetti F, Gu S, Cieslak M, Grafton ST, Vettel JM, et al. Stimulation-based control of dynamic brain networks. arXiv. 2016;In Prep. 26. Ravizza SM, Moua KC, Long D, Carter CS. The impact of context processing deficits on task-switching performance in schizophrenia. Schizophr Res. 2010;116(2 -- 3):274 -- 279. 27. Wylie GR, Clark EA, Butler PD, Javitt DC. Schizophrenia patients show task switching deficits consistent with N-methyl-d-aspartate system dysfunction but not global executive deficits: implications for pathophysiology of executive dysfunction in schizophrenia. Schizophr Bull. 2010;36(3):585 -- 594. 28. Cieslak M, Grafton ST. Local termination pattern analysis: a tool for comparing white matter morphology. Brain Imaging Behav. 2014;8(2):292 -- 299. 29. Yeh FC, Tseng W. NTU-90: a high angular resolution brain atlas constructed by q-space diffeomorphic reconstruction. Neuroimage. 2011;58(1):91 -- 99. 30. Dale AM, Fischl B, Sereno MI. Cortical surface-based analysis. I. Segmentation and surface reconstruction. NeuroImage. 1999;9(2):179 -- 194. 31. Hagmann P, Cammoun L, Gigandet X, Meuli R, Honey CJ, Wedeen VJ, et al. Mapping the structural core of human cerebral cortex. PLoS Biology. 2008;6(7):e159. 32. Bassett DS, Brown JA, Deshpande V, Carlson JM, Grafton ST. Conserved and variable architecture of human white matter connectivity. Neuroimage. 2011;54(2):1262 -- 1279. 33. Hermundstad AM, Bassett DS, Brown KS, Aminoff EM, Clewett D, Freeman S, et al. Structural foundations of resting-state and task-based functional connectivity in the human brain. Proc Natl Acad Sci U S A. 2013;110(15):6169 -- 6174. PLOS 18/21 34. Hermundstad AM, Brown KS, Bassett DS, Aminoff EM, Frithsen A, Johnson A, et al. Structurally-constrained relationships between cognitive states in the human brain. PLoS Comput Biol. 2014;10(5):e1003591. 35. Biol PC. Resolving structural variability in network models and the brain. 10;(3):e1003491. 36. Small-World Propensity in Weighted, Real-World Networks. arXiv. 2015;1505:02194. 37. Sizemore A, Giusti C, Bassett DS. Classification of weighted networks through mesoscale homological features. arXiv. 2015;1512:06457. 38. Yeh FC, Wedeen VJ, Tseng WY. Generalized Q-Sampling Imaging. Medical Imaging, IEEE Transactions on. 2010;29(9):1626 -- 1635. 39. Tuch DS. Q-ball imaging. Magnetic Resonance in Medicine. 2004;52(6):1358 -- 1372. 40. Baird B, Smallwood J, Gorgolewski KJ, Margulies DS. Medial and lateral networks in anterior prefrontal cortex support metacognitive ability for memory and perception. The Journal of Neuroscience. 2013;33(42):16657 -- 16665. 41. Satterthwaite TD, Elliott MA, Gerraty RT, Ruparel K, Loughead J, Calkins ME, et al. An improved framework for confound regression and filtering for control of motion artifact in the preprocessing of resting-state functional connectivity data. Neuroimage. 2013;64:240 -- 256. 42. Watanabe T, Hirose S, Wada H, Imai Y, Machida T, Shirouzu I, et al. A pairwise maximum entropy model accurately describes resting-state human brain networks. Nat Commun. 2013;4:1370. 43. Watanabe T, Hirose S, Wada H, Imai Y, Machida T, Shirouzu I, et al. Energy landscapes of resting-state brain networks. Front Neuroinform. 2014;8:12. 44. Newman MEJ. Modularity and community structure in networks. Proceedings of the National Academy of Sciences of the United States of America. 2006;103(23):8577 -- 8696. 45. Clauset A, Newman ME, Moore C. Finding community structure in very large networks. Physical review E. 2004;70(6):066111. 46. Cipra BA. The Ising model is NP-complete. SIAM News. 2000;33(6):1 -- 3. 47. Metropolis N, Rosenbluth AW, Rosenbluth MN, Teller AH, Teller E. Equation of state calculations by fast computing machines. The journal of chemical physics. 1953;21(6):1087 -- 1092. 48. Manning CD, Raghavan P, Schutze H, et al. Introduction to information retrieval. vol. 1. Cambridge university press Cambridge; 2008. 49. Power JD, Cohen AL, Nelson SM, Wig GS, Barnes KA, Church JA, et al. Functional network organization of the human brain. Neuron. 2011;72(4):665 -- 678. 50. Evolution of brain network dynamics in neurodevelopment. arXiv. 2016;. 51. Clauset A, Shalizi CR, Newman MEJ. Power-law distributions in empirical data. SIAM Review. 2009;51(4):661 -- 703. PLOS 19/21 52. Y V, Clauset A. Power-law distributions in binned empirical data. Annals of Applied Statistics. 2014;8(1):89 -- 119. 53. Crossley NA, Mechelli A, Vertes PE, Winton-Brown TT, Patel AX, Ginestet CE, et al. Cognitive relevance of the community structure of the human brain functional coactivation network. Proc Natl Acad Sci U S A. 2013;110(28):11583 -- 11588. 54. Crossley NA, Mechelli A, Scott J, Carletti F, Fox PT, McGuire P, et al. The hubs of the human connectome are generally implicated in the anatomy of brain disorders. Brain. 2014;137(Pt 8):2382 -- 2395. 55. Raichle ME, MacLeod AM, Snyder AZ, Powers WJ, Gusnard DA, Shulman GL. A default mode of brain function. Proceedings of the National Academy of Sciences. 2001;98(2):676 -- 682. 56. Sporns O, Tononi G, Kotter R. The human connectome: A structural description of the human brain. PLoS Comput Biol. 2005;1(4):e42. 57. Sporns O. The human connectome: a complex network. Ann N Y Acad Sci. 2011;1224:109 -- 125. 58. Bassett DS, Yang M, Wymbs NF, Grafton ST. Learning-induced autonomy of sensorimotor systems. Nat Neurosci. 2015;18(5):744 -- 751. 59. Sporns O. Cerebral cartography and connectomics. Philos Trans R Soc Lond B Biol Sci. 2015;370(1668):20140173. 60. Zeki S. Introduction: cerebral cartography 1905 -- 2005. Phil Trans R Soc B. 2005;360:651 -- -652. 61. Newman MEJ, Clauset A. Structure and inference in annotated networks. arXiv. 2015;1507:04001. 62. Schirner M, Rothmeier S, Jirsa VK, McIntosh AR, Ritter P. An automated pipeline for constructing personalized virtual brains from multimodal neuroimaging data. Neuroimage. 2015;117:343 -- 357. 63. Ritter P, Schirner M, McIntosh AR, Jirsa VK. The virtual brain integrates computational modeling and multimodal neuroimaging. Brain Connect. 2013;3(2):121 -- 145. 64. Bassett DS, Nelson BG, Mueller BA, Camchong J, Lim KO. Altered resting state complexity in schizophrenia. Neuroimage. 2012;59(3):2196 -- 2207. 65. Zalesky A, Fornito A, Egan GF, Pantelis C, Bullmore ET. The relationship between regional and inter-regional functional connectivity deficits in schizophrenia. Hum Brain Mapp. 2012;33(11):2535 -- 2549. 66. Yu Q, Sui J, Liu J, Plis SM, Kiehl KA, Pearlson G, et al. Disrupted correlation between low frequency power and connectivity strength of resting state brain networks in schizophrenia. Schizophr Res. 2013;143(1):165 -- 171. 67. Honey CJ, Sporns O, Cammoun L, Gigandet X, Thiran JP, Meuli R, et al. Predicting human resting-state functional connectivity from structural connectivity. Proc Natl Acad Sci U S A. 2009;106(6):2035 -- 2040. PLOS 20/21 68. Becker C, Pequito S, Pappas GJ, Miller MB, Grafton ST, Preciado V. Accurately Predicting Functional Connectivity from Diffusion Imaging. arXiv. 2016;1512:02602. 69. Cole MW, Bassett DS, Power JD, Braver TS, Petersen SE. Intrinsic and task-evoked network architectures of the human brain. Neuron. 2014;83(1):238 -- 251. 70. Bialek W, Cavagna A, Giardina I, Mora T, Silvestri E, Viale M, et al. Statistical mechanics for natural flocks of birds. Proc Natl Acad Sci U S A. 2012;109(13):4786 -- 4791. 71. Giusti C, Ghrist R, Bassett DS. Two's company, three (or more) is a simplex: Algebraic-topological tools for understanding higher-order structure in neural data. Trends in Cognitive Sciences. 2016;Under Consideration. 72. Leen DA, Shea-Brown E. A Simple Mechanism for Beyond-Pairwise Correlations in Integrate-and-Fire Neurons. J Math Neurosci. 2015;5(1):30. 73. Fox PT, Lancaster JL, Laird AR, Eickhoff SB. Meta-analysis in human neuroimaging: computational modeling of large-scale databases. Annu Rev Neurosci. 2014;37:409 -- 434. 74. Pasqualetti F, Zampieri S, Bullo F. Controllability Metrics, Limitations and Algorithms for Complex Networks. IEEE Transactions on Control of Network Systems. 2014;1(1). 75. Wu S, Hitchman G, Tan J, Zhao Y, Tang D, Wang L, et al. The neural dynamic mechanisms of asymmetric switch costs in a combined Stroop-task-switching paradigm. Sci Rep. 2015;5:10240. 76. Davidson MC, Amso D, Anderson LC, Diamond A. Development of cognitive control and executive functions from 4 to 13 years: evidence from manipulations of memory, inhibition, and task switching. Neuropsychologia. 2006;44(11):2037 -- 2078. PLOS 21/21
1904.06712
1
1904
2019-04-14T15:46:37
Bilateral symmetry strengthens the perceptual salience of figure against ground
[ "q-bio.NC" ]
Although symmetry has been discussed in terms of a major law of perceptual organization since the early conceptual efforts of the Gestalt school (Wertheimer, Metzger, Koffka and others), the first quantitative measurements testing for effects of symmetry on processes of Gestalt formation have seen the day only recently. In this study, a psychophysical rating study and a "foreground" versus "background" choice response time experiment were run with human observers to test for effects of bilateral symmetry on the perceived strength of figure against ground in triangular Kanizsa configurations. Displays with and without bilateral symmetry, identical physically specified to total contour ratio and constant local contrast intensity within and across conditions, but variable local contrast polarity and variable orientation in the plane were presented in a random order to human observers. Configurations with bilateral symmetry produced significantly stronger figure against ground percepts reflected by greater subjective magnitudes and consistently higher percentages of "foreground" judgments accompanied by significantly shorter response times. These effects of symmetry depend neither on the orientation of the axis of symmetry, nor on the contrast polarity of the physical inducers. It is concluded that bilateral symmetry, irrespective of orientation, significantly contributes to the, largely sign invariant, visual mechanisms of shape segregation that determine the salience of figure against ground in perceptually ambiguous image configurations.
q-bio.NC
q-bio
Bilateral symmetry strengthens the perceptual salience of figure against ground Birgitta Dresp-Langley 1* 1 Centre National de la Recherche Scientifique, ICube Lab UMR 7357 CNRS and Strasbourg University, Strasbourg, FRANCE; [email protected] * Correspondence: [email protected] configurations. Displays with Abstract: Although symmetry has been discussed in terms of a major law of perceptual organization since the early conceptual efforts of the Gestalt school (Wertheimer, Metzger, Koffka and others), the first quantitative measurements testing for effects of symmetry on processes of Gestalt formation have seen the day only recently. In this study, a psychophysical rating study and a "foreground"-"background" choice response time experiment were run with human observers to test for effects of bilateral symmetry on the perceived strength of figure-ground in triangular Kanizsa identical physically-specified-to-total contour ratio and constant local contrast intensity within and across conditions, but variable local contrast polarity and variable orientation in the plane were presented in a random order to human observers. Configurations with bilateral symmetry produced significantly stronger figure-ground percepts reflected by greater subjective magnitudes and consistently higher percentages of "foreground" judgments accompanied by significantly shorter response times. These effects of symmetry depend neither on the orientation of the axis of symmetry, nor on the contrast polarity of the physical inducers. It is concluded that bilateral symmetry, irrespective of orientation, significantly contributes to the, largely sign-invariant, visual mechanisms of figure-ground segregation that determine the salience of figure-ground in perceptually ambiguous configurations. and without bilateral symmetry, Keywords: bilateral symmetry; physically-specified-to-total contour ratio; Kanizsa's triangle; figure-ground segregation; law of good Gestalt; subjective magnitude estimation; choice response time 1. Introduction The Gestalt psychologists Max Wertheimer and Wolfgang Metzger [1, 2] formulated and discussed "laws of perception" to predict how perceptual grouping operates under specific conditions of visual configuration. Their important work was translated into the English language in 2012 and 2009 respectively by Lothar Spillmann and colleagues [1, 2], making this important early conceptual work available to a broader audience. In physical science, a law is a prediction that can be proven true and, ideally, the limits of which can also be clearly determined. In perceptual science, the Gestalt laws are used to express principles or conditions of visual configuration to explain why we see the world as we do. It is argued that specific principles of, or conditions for, "good Gestalt" need to be fulfilled to enable what is called perceptual grouping, i.e. a perceptual solution that yields the most plausible interpretation of a given physical configuration. Since all physical stimuli are by nature ambiguous to our perception, they need to be interpreted by the brain to produce coherent and unambiguous percepts that allow us to act on the physical world effectively. The "Law of Symmetry" is a major Gestalt law. It predicts that visual elements that are symmetrical would be more likely to form a group, i.e. to be perceived as a "good Gestalt", in comparison with asymmetrical ones. Visual symmetry has, indeed, proven a determining factor in shape perception [3-6]. In particular, vertical mirror symmetry has proven an important cue to shape extraction from abstract, non-familiar visual elements presented in conditions of heightened ambiguïty (noise). Across different noise levels, symmetric elements form perceptually more salient shapes than asymmetric ones and are, therefore, more readily detected [5]. The Italian Gestalt psychologist Gaetano Kanizsa [7] discussed a series of ambiguous planar configurations that give rise to powerful figure-ground percepts, with apparent shapes emerging in the foreground, delineated by contours that are perceptually completed beyond physically specified contrast edges. The Gestalt school and Kanizsa himself considered these phenomena as marginal cases of perception ("margini quasi-percettivi"), and argued that these latter provide insight into the fundamentals of perceptual organization because they put underlying processes to the test under extreme conditions at the capacity limits of the perceptual system. Later-on, the figures seen in such configurations were termed "illusory" by cognitive psychologists; the Gestalt psychologists themselves never used this term, which is, of course, misleading. An illusion, by definition, cannot be verified by independent observation - it only exists in the mind of the person experiencing it. The phenomena described by Kanizsa have clearly defined physical correlates, with measurable systematic effects on perception. One of these configurations is the famous Kanizsa triangle (Figure 1). The Kanizsa figures have been studied extensively to single out factors of physical variation that affect the subjective brightness or darkness of the figures and/or the figure contours. The results from these studies, based on a variety of different experimental measures, are reviewed in sufficient detail elsewhere [8-13]. They are not the object of this study here. Here, we measured the perceptual salience of the figure-ground percept irrespective of the relative darkness, brightness, or clarity of either the induced surfaces or their boundaries, as is made perfectly clear in the instructions given to subjects. As raised previously by others, the response criterion of the subjects in judgment tasks using this type of ambiguous figure here [8] is directly dependent on the semantic precision of instructions given. Formulating these latter appropriately is to make sure the perceptual phenomenon under study, not a related one, is reflected by the psychophysical data. The influence of variations in the intensity of local luminance contrast of the physical inducers that produce a perceptual filling in [14] of bright or dark surfaces, leading to figure-ground percepts at the centre of configurations, was demonstrated for the first time by Heinemann's pioneering studies [15]. These were extended later on by others [12] to various configurations of the Kanizsa type, where observers had to adjust the luminance of the central figure region until it matched the phenomenal appearance of the general background. This is equivalent to a cancellation of the phenomenal appearance of figure and ground. The configurations produce filling-in consistent with classic simultaneous contrast effects where the central figure appears darker than the general background when surrounded by phenomenally white inducers, and brighter when surrounded by phenomenally black inducers. The perceptual salience of the central figures increases consistently with the luminance contrast intensity of the inducers, up to some optimal limit. When that optimal limit is reached and the contrast intensity of inducers increases further, the figure intensity is diminished again and may be annulled completely at the highest physical contrast levels [15]. The simultaneous contrast filling-in that leads to the figure-ground percepts in Kanizsa configurations is thus predicted by specific physical parameters of the inducers, which were all controlled experimentally to keep them constant across symmetry conditions in this study here. When the physical contrast intensity of the configurational elements is optimal [12-15] and not varied between displays, the next most important physical parameter that straightforwardly determines the salience or subjective strength of the figure-ground percept in Kanizsa configurations is the physically specified-to-total contour ratio, or support ratio. This was proven in a series of experiments by Shipley and Kellmann [11] using subjective magnitude estimation, a classic psychophysical rating procedure similar to the one applied in this study. Here, the Kanizsa triangle is exploited to probe for effects of symmetry on figure-ground from occlusion cues. The Kanizsa triangle is one of the most cited examples of a specific class of Gestalt configurations where perceived surface depth arises from the local occlusion cues. In this specific shape class, figure-ground results directly from a process of surface completion through boundary interpolation across the physically specified edge contours of the inducers providing the occlusion cues [16-20]. Occluded object completion thus reflects the workings of fundamental visual mechanisms for recovering object percepts from fragmented input, and the ability of human perception to read structure into an apparently chaotic physical world [21, 22]. The functional interactions between configurational symmetry and other structural factors in this important perceptual process are still unknown. The dependency of figure-ground salience on the support ratio (Figure 1), a scale invariant metric, is associated with the ecologically desirable consequence that perceptual salience will not change with variations in viewing distance [11]. At constant physical contrast intensity of configurations with a constant support ratio, the contrast polarity of inducers, i.e. whether they are dark on lighter backgrounds, bright on darker backgrounds or a mixture of both on a medium grey background, does not affect the salience or subjective strength of the resulting figure-ground percept, provided the contrast polarity is homogenous within each configurational element [12, 23, 24, 25, 28, 29]. When contrast polarity is not homogenous within the configurational elements, then, and only then, may the perceptual salience of the figure-ground percept be reduced [26], or not [8] depending on task instructions. Perceptual figure -- ground organization is thus determined by visual mechanisms that integrate the contrast intensity and the spatial information carried by the configurational elements while mostly discarding information on contrast polarity. This is predicted by sign-invariant models based on the functional properties of cortical neurons of the complex type, which are orientation selective but insensitive to contrast polarity [14, 23, 24, 25, 29, 35]. Such sign-invariant visual mechanisms have the ecologically desirable consequence that the simplest plausible representation of figure and ground can be achieved when the signal input from local contrast regions is ambiguous. The classic version of the Kanizsa triangle is a configuration with perfect vertical bilateral, or mirror symmetry (Figure 1, top, and bottom left). Whether physical display variations producing asymmetric configurations (Figure 1, bottom right) would affect figure-ground salience in this specific case is not known. The motivation of this study was, therefore, to test whether symmetry contributes to figure-ground strength in one of the most classic Gestalt configurations where surface depth results from visual interpolation across fragments. Two variations of the Kanizsa triangle with identical support ratio as defined by Shipley and Kellmann [11] and identical triangular area size were generated; one with perfect bilateral symmetry (Figure 1, bottom left), the other asymmetric (Figure 1, bottom right). To test for possible interactions between symmetry and the orientation of the configurations in the plane, presentations were varied and bilateral symmetry was not always vertical but could be vertical or horizontal, in a random order. In the light of earlier findings, with abstract shapes presented in noisy contexts [5], vertical mirror symmetry significantly increased the probability that a shape was seen as figure against the ground. Thus, the bilateral symmetry of vertical orientations may also generate stronger effects on the figure-ground salience of surfaces completed by interpolation. The physical inducers, either dark on grey, light on grey, or light and dark on grey, displayed variations in contrast polarity across inducers, but never within, in both types of configuration, symmetric and asymmetric. In the light of previous findings, these variations should not affect figure-ground strength, given that the polarity of contrast was always homogenous within inducers in the different configurations [8, 10, 12, 26, 29]. In a first experiment, psychophysical magnitude estimation was used to measure the salience, or subjective strength, of the figure-ground percepts in the different conditions. In a second experiment, a choice response time was run, with a selected set of configurations, to test whether more salient figure-ground percepts consistently produce, as would be expected, shorter response times with consistent "foreground" response probabilities. Figure 1. Variations of the Kanizsa triangle with phenomenally black inducers on a grey background. The subjective strength of the triangular figure-ground percept emerging in the center of the configuration critically depends on the physically specified-to-total contour, or support ratio. Stronger surface percepts are produced by higher support ratios (top). To test for effects of symmetry on the salience of figure-ground, configurations with bilateral symmetry (bottom left) and without (bottom right) were generated. All the configurations had identical support ratio and therefore identical area size. All physically specified elements in the configurations (inducers) were of identical size and contrast intensity. 2. Materials and Methods Triangular Kanizsa configurations with and without bilateral symmetry (Figure 1 bottom left and right respectively, for illustration), identical support ratio and surface area, variable orientation (vertical base down, as in Figure 1 bottom left, vertical base up, sideways base left, sideways base right) and variable inducer polarity (three black inducers on grey or '- - -', three inducers white on grey or '+++', two black and one white inducer on grey or '- - +', two white and one black inducer on grey or '++ -') were presented in random order to ten human observers in a single-presentation-per- figure subjective magnitude estimation (rating) experiment with moduli. Four of these configurations with a single orientation (vertical base down, as in Figure 1 bottom left) and uniformly positive (+++) or uniformly negative (---) contrast polarity, two with vertical symmetry and two without, were presented to six additional human observers in a repeated measures (four presentations per configuration) choice response time experiment with two additional control stimuli (triangles with minimally visible line contours ("ghosts") only, no surface contrast). Stimuli The image configurations were computer generated using an HP Zbook 15 G2 Mobile Workstation equipped with a 4th generation Intel Core i7-6700 processor and an NVIDIA Quadro K5100 graphic card. Table 1. Figure dimensions in centimeters (cm) with the overall support ratio and surface area as a function of configuration (symmetric versus asymmetric). The symmetry factor only varies systematically between configurations, the shape interpretation ("triangle") is the same and so are all relevant physical parameters. Symmetric Asymmetric Triangle base (b) 9 cm 13 cm Triangle side 1 Triangle side 2 Triangle height (h) Triangle surface area (1/2bxh) Physical inducer radius Support ratio 12 cm 12 cm 11 cm 49.5 cm 2 cm 0.36 11 cm 9 cm 7.62 cm 49.5 cm 2 cm 0.36 Configurational dimensions in terms of size (in pixels and in cm on screen) of triangle base and triangle sides, the inducer radius, which was identical in all the configurations, and the overall physical-to-total contour ratio, also identical in all the configurations, are summarized in Table 1. Luminance values of the different configural elements were determined photometrically using an OPTICAL photometer (Cambridge Research Systems). The ADOBE RGB coordinates of the phenomenally grey background (RGB: 140, 140, 140) yield a background luminance of 55 cd/m2. The phenomenally black inducers (RGB: 5, 5, 5) a luminance of 4 cd/m2, and the phenomenally white inducers (RGB: 240, 240, 240) a luminance of 98 cd/m2. The moduli from the sujective rating task (RGB: 135, 135, 135 for the phenomenally darker ones and RGB: 145, 145, 145 for the phenomenally lighter ones) had a luminance of 52 cd/m2 or 58 cd/m2 respectively. The line contour control configurations (RGB: 120, 120, 120) from the choice response time task had a luminance of 48 cd/m2. The physically specified contrast intensities with positive and negative signs may be calculated using the Weber Contrast (Weber Ratio, W) formula: W = (Lconfig-Lbackground)/Lbackground (1). As a consequence, we have a positive W of +0.92 for the phenomenally white inducers, a negative W of -.78 for the phenomenally black inducers, a positive and a negative W of +.09 and -.09 respectively for the minimal-contrast moduli from the subjective rating task, and a negative W of -.13 for the minimal contrast line contour control configurations from the choice response time experiment. Presentation of configurations The configurations were presented in random order on the screen of the HP Zbook 15 G2 Mobile Workstation, which has a pixel resolution of 1920x1080 and a 60 Hz refresh rate. Random selection, presentation, and response coding were computer controlled using Python for Windows. The duration of presentation of each single configuration was observer controlled in both experiments, a subsequent presentation always initiated 800 milliseconds after the observer had typed his/her response on the single-presentation subjective magnitude estimation (rating) task, with the different variations in orientation and in local contrast polarity, are shown in Figure 2 a and b, for illustration. Illustrations of the 6 configurations from the repeated measures choice response time task are shown in Figure 3. the computer keyboard. The 32 configurations from Experimental procedure Subjects were seated in front of the workstation at a distance of about 90 cm from the screen in a semi-dark room. In the subjective magnitude estimation task, they were shown a set of moduli consisting of minimal-contrast triangular surfaces of the same spatial dimensions as the symmetric and asymmetric triangular centers of the test configurations. These moduli are shown in Figure 2c for illustration. Subjects were told to associate the phenomenal strength of the moduli with a rating score of '11'. It was then made clear to them that they would be shown different triangular configurations, with black and white patches around them. They were then asked to rate the subjective strength of the figure-ground percept at the centre of the test configurations, in terms of the strength of the segregation into foreground and background, regardless of the direction of the perceived contrast (i. e. subjectively darker or subjectively lighter), by a number between '0' and '10', bearing in mind that the highest number was to reflect a figure salience closest to that of the real-contrast moduli they had seen just before. Each of the 16 asymmetric and the 16 symmetric configurations (Figure 2a and b, respectively, for illustration) was presented only once to each of the subjects in a single random order session. In the choice response time task, subjects were asked to judge as swiftly as possible whether the triangle displayed on the screen seemed to stand out as foreground against the grey general background, or to lie behind the general grey background. In this experiment, the outlined triangular shape control configurations (Figure 3, on the right) were presented at a minimal, just visible negative line contrast intensity and no surface contrast. This renders a highly ambiguous (one subject mentioned "ghost-like") appearance on the screen with no clear figure-ground assignment. Each of the six configurations (Figure 3, for illustration) was shown four times, in random order, to each of six subjects in a single individual session. Subjects Ten individuals (six men, four women) between 20 and 31 years old, all of them with normal or corrected to normal vision, participated in the subjective magnitude estimation experiment. Six further individuals (five men, one woman), also young and with normal or corrected to normal vision, participated time experiment. Participants were mostly undergraduates involved in medical or language studies. None of them was familiar with the configurations presented to them, and all of them were naïve to the purpose of the study. The experiments were conducted in accordance with the Declaration of Helsinki (1964) and in full conformity with the author's host institution's (CNRS) ethical standards committee. Informed consent was obtained from each of the participants. the choice response in a) b) c) Figure 2. The configurations from the subjective rating experiment (a) The asymmetric Kanizsa triangles, with varying orientation and inducer contrast polarities (b) The Kanizsa triangles with axial symmetry c) the moduli for benchmarking the subjective rating scale (0-10). Subjects were told to associate the moduli with a figure-ground strength rating of '11', regardless of the direction of the perceived contrast. Figure 3. The test and control configurations ("ghosts") from the choice response time experiment. Six subjects were asked to judge as swiftly as possible whether the triangle displayed on the screen seemed to stand out as foreground against the grey general background, or to lie behind the general grey background. Data analysis The data from the subjective rating experiment, with a Cartesian design plan written in terms of Subject10xSymmetry2xOrientation4xPolarity4, produced a total of 320 subjective ratings. These data were fed into a Three-Way ANOVA. Means, standard errors, effect sizes, and F statistics with probability limits were determined. The data from the choice response time experiment, with a Cartesian design plan written in terms of Subject6xSymmetry2xPolarity3xRepeatedMeasures4, produced a total of 144 choice data and a total of 144 response times. In the experimental design plan, the control configuration represents the third modality of the "polarity" factor, with the three factor levels "positive" or '+++', "negative" or '- - -', and "control". The response times were fed into a Two-Way Repeated Measures ANOVA with individual data averaged over the four levels of the repetition factor R4 and without the third level of the "polarity" factor, i.e. the analysis plan therefore reads Subject6xSymmetry2xPolarity2. 3. Results The results of the analyses on the data from the subjective rating experiment and the choice response time experiment are shown here below in the form of graphs and tables. 3.1. Subjective magnitude estimation task The individual data from this experiment are made available in Table S1 of the Supplementary Materials section. The data show a good consistency between participants, within and across conditions, with a moderate amount of inter-individual variability. The main effect of symmetry on the subjective magnitude of the figure-ground percept in the Kanizsa configuration is highlighted by the two graphs in Figure 4. The average magnitude of figure salience in terms of average subjective ratings produced by symmetric and asymmetric configurations is plotted as a function of the orientation of the configurations in the plane (Figure 4, top), and as a function of the contrast polarity of the inducers (Figure 4, bottom). While the subjective ratings display no major variations with either orientation of the configurations or the contrast polarity of the inducers, they are consistently and noticeably affected by lack of symmetry. Subjective ratings are found to be markedly stronger for all the configurations with bilateral symmetry, irrespective of orientation and/or contrast polarity. This effect is highlighted further by the statistics shown in Table 2 here below, which summarizes the observations from Figure 4 in terms of results from the Three-Way ANOVA on the subjective ratings of the ten subjects. The effect of symmetry is statistically significant, the effects of orientation and inducer polarity are not. Table 2. Three-Way ANOVA results with the means (average subjective magnitudes), standard errors (SEM), and F statistics for effects of main factors and their interactions from the analysis of the subjective rating data Level Mean SEM F 3.8 6.1 4.7 5.3 4.8 4.7 4.8 5.2 4.9 4.7 _ _ _ F(3, 319)=1.24; NS 0.15 F (1, 319)=108.8; p<.001 0.13 0.21 0.22 0.20 0.19 0.20 0.23 0.21 0.22 F(3, 319)=1.17; NS _ _ _ F(3, 319)=1.83; NS F(3, 319)=0.41; NS F(9, 319)=0.67; NS Factor Symmetry (S2) Polarity (P4) asymmetric symmetric - - - +++ - - + ++ - Orientation (O4) vertical base bottom vertical base top sideways base left sideways base right Symmetry x Polarity Symmetry x Orientation Polarity x Orientation interaction interaction interaction Choice response time task The raw data from this experiment are made available in Table S2 of the Supplementary Materials section. The perceptual judgments from the choice task, shown here below in Table 3 in terms of the percentage of "foreground" responses as a function of configuration (asymmetric vs symmetric) and the contrast polarity of the inducers (negative vs positive vs control), expressed in terms of the phenomenal appearance of the inducers here, display a consistent majority of "foreground" responses in the ambiguous Kanizsa configurations, with noticeably higher percentages of "foreground" in the configurations with bilateral symmetry, irrespective of inducer appearance (or polarity). In response to the control configurations, the percentages of "foreground" responses show no such clear trend. This is explained by the fact that the outlined shape control configurations (Figure 3, on the right) were presented at a minimal, just visible negative line contrast intensity, which made them particularly ambiguous with respect to figure-ground organization in the plane. The outlines are not perceived as clearly belonging to a specific depth level, which is reflected in the results here by a near random distribution of "foreground" and "background" responses. This suggests that the outlined controls without surface contrast did not produce, as could be expected, salient figure-ground percepts. Average choice response times were plotted as a function of configuration and inducer polarity, as shown here in Figure 5. The graphs show a consistent and systematic effect of symmetry on response times. Subjects respond markedly faster to configurations with axial symmetry, irrespective of whether the inducers are phenomenally black (negative contrast polarity) or white (positive contrast polarity). This effect is highlighted further by the statistics shown in Table 4 here below, which summarizes the observations from Figure 5 in terms of results from the Two-Way Repeated Measures ANOVA on the response times of the six subjects. The effect of symmetry is statistically significant, the effect of inducer polarity is not. The third level of the "polarity" factor here, i.e. the control configuration, was not included in the design plan for this ANOVA. Figure 4. Average magnitudes of figure-ground in terms of average subjective ratings with bars indicating +/- the standard error of the mean. Effects produced by symmetric and asymmetric configurations are plotted as a function of the orientation of configurations in the plane (top), and as a function of inducer polarity (bottom). Table 3. Percentage of "foreground" responses from the choice response time task as a function of configuration and inducer contrast polarity Asymmetric Symmetric White inducers Black inducers Control 88 % 75 % 70 % 98 % 92 % 55 % Figure 5. Average choice response times with bars indicating +/- the standard error of the mean as a function of configuration and inducer polarity. Table 4. Two-Way Repeated Measures ANOVA results with the means (in milliseconds), standard errors (SEM), and F statistics for effects of main factors and their interactions from the analysis of the choice response times Factor Symmetry (S2) Polarity (P2) Level asymmetric symmetric - - - +++ Symmetry x Polarity interaction _ 4. Discussion Mean SEM 1518 900 1269 1150 73 65 64 62 _ F F(1, 23)=36.69; p<.01 F(1, 23)=1.74; NS F(1, 23)=1.21; NS Although symmetry has been discussed in terms of a major grouping principle or law of good Gestalt since Wertheimer and Metzger [1, 2], the specific effects symmetry may produce on feature grouping, figure-ground segregation, visual discrimination, or time to respond to visual configurations have become subject to systematic quantitative investigation in perceptual science only recently. In the case of visual perception, symmetry may be conceived as a geometric property that yields configurational simplicity and, therefore, represents an ecological advantage for information processing [30]. Symmetry may also be seen as a perceptual feature that attracts attention, enhances configural salience, and facilitates grouping [5, 31-33]. It is found that bilateral configurational symmetry, i.e. mirror symmetry within the whole configuration, strengthens the perceptual salience of figure against ground in triangular Kanizsa configurations. The results from the magnitude estimation (rating) experiment clearly show that the subjective strength of the foreground at the center of the configurations is significantly higher when the configurations have bilateral symmetry. This holds for triangular configurations when mirror symmetric configurations are compared to asymmetric configurations with the same number of inducers, and the same support ratio as defined by Shipley and Kellman [11]. This symmetry effect can be exploited to further quantify critical interactions between occlusion-based surface properties, symmetry, and figure-ground salience. Variations in symmetry could be tested against variations in the support ratio in a first instance. Figure-ground segregation from interpolation is an early-stage process in perceptual grouping [16, 44, 45], as is symmetry detection [39, 43]. In particular, as shown by Erlikhman and Kellmann [44, 45], the human perceptual system uses critical spatial cues of local position and alignment within a restricted spatiotemporal window (~165 msec) for the rapid extraction of co-oriented edge fragments from the visual input. As predicted by the Gestalt Law of Good Continuation, the fragments then connect by known neural interpolation processes [16, 25, 35], producing larger surfaces that will stand out as figures against ground. The results from the experiments here show that symmetry contributes to this early process of perceptual organization. The results suggest no influence of the orientation (vertical versus horizontal) of the axis of symmetry on the salience of the figure-ground percepts. Vertical and horizontal mirror symmetry produced equally strong phenomenal salience of figure-ground. Since Mach [33], it is suggested that symmetry around the vertical axis may be more effectively processed by the visual system than symmetry about the horizontal, or any other, axis in the plane. Some studies have confirmed this prediction [5]. However, more recent reviews indicate that there may be no systematic functional advantage of vertical symmetry [34]. Effects of axis of symmetry on perception may be dependent on what Bertamini termed "objectness" [31], i.e. whether the cognitive interpretation of the visual shape changes with translational or rotational changes of the latter. Psychophysical data on shape perception [43], using radial frequency patterns and other objects, indeed suggest that variations in the location and orientation of relevant (with respect to the perceptual task) object features may generate effects of axis of symmetry. In the two experiments here, relevant perceptual features within and across objects (symmetric verus asymmetric) can be considered invariant, since there was no effect of contrast polarity and no interaction between contrast polarity and symmetry. This could explain why the axis of symmetry had no effect here either. Also, earlier psychophysical studies have shown that bilateral symmetry is significantly more salient within objects, significantly less between objects [39, 40]. The layout characteristics, including symmetry, of complex figure-ground solutions are more easily processed within single perceptual objects [40]. Symmetry detection becomes harder with complex shape configurations where other factors, such as positional uncertainty or convexity, interact with the symmetry factor, especially when the psychophysical task requires comparing across objects. It may be that vertical symmetry generates a measurable advantage for perception only under such conditions. Figure 6. An example of a configuration where effects of symmetry on visual perception cannot be tested independently from possible effects of shape interpretation. In the square version of the Kanizsa figure, breaking the symmetry of the classic square configuration (images on left) inevitably requires breaking the perpendicularity of the shape borders. This inevitably results in a new and qualitatively different shape geometry and shape interpretation. In other words, shape interpretation then becomes a confusion factor. In this case here a "shard" with qualitatively different 3D-like shape properties emerges (images in middle and on left). When shape orientation changes, the percept changes, again, qualitatively and a new visual object emerges (cf. on the importance of "objectness", see Bertamini [31]). All configurations here above have roughly equivalent, albeit not strictly identical, support ratio and central area size. Variations in inducer texture, figure orientation, and background intensity are presented here for illustration only. The results from the choice response time task show a consistently higher percentage of "foreground" responses to the symmetric configurations, which is accompanied by significantly shorter response times. Bilateral symmetry, therefore, represents a measurable functional advantage in the perceptual processing of figure-ground. Variations in the contrast polarity of the inducing elements had no marked effect on either the subjective strength of the figure-ground percepts or the response times. This observation is consistent with results from earlier work with similar configurations where symmetry was not varied [9, 12, 29], and predicted by non-linear neural models of figure-ground based on the long-range integration of antagonistic brightness signals [12, 13, 14, 23, 24]. Interestingly, when inducers of both positive and negative contrast signs, i.e. phenomenally white and black inducers, are present in the same configuration, the latter may be perceived as phenomenally asymmetric with respect to brightness. The perceptual system, however, is not influenced by the symmetry/asymmetry in contrast signals, only by geometrically defined symmetry. Although this study here was not specifically aimed at singling out the hierarchal level of perceptual processing at which the symmetry effect arises, it is unlikely that conscious processing was involved here. After the experiments, subjects were asked whether they had noticed anything in particular in the configurations or used a particular strategy to respond. Most of them stated that "some were the same, some were different", or "some had white parts, some had black parts, some had both", but none of them was able to specify any particular structural difference or response strategy. We may therefore infer that subjects were not immediately aware of the systematic variations in symmetry. Bilateral symmetry is identified as a key prior for three-dimensional shape perception in humans [41, 42]. The perceptual integration of symmetry in this process does not necessarily happen consciously and, as explained above, may vary with shape complexity and shape interpretation [40] without subjects being aware of it. Therefore, configurational complexity and shape interpretation need to be controlled for to single out the effects of symmetry per se on any particular aspect of perceptual organization. This is clarified further by some of the additional illustrations in Figure 6, showing configurations where the manipulation of symmetry inevitably implies changing also the complexity of the configuration as a whole, and the resulting shape interpretation. In the square version of the Kanizsa figure, breaking the configurational symmetry inevitably requires changing the shape borders. This produces a new, far more complex, qualitatively different shape geometry leading to a radically different shape interpretation. The Kanizsa square is therefore ill-suited for singling out effects of symmetry without ambivalence. The advantage of the triangular configuration used in this study here is that the geometric transformations needed to manipulate mirror symmetry affect neither the structural complexity of the configurations, nor the resulting shape interpretation: with or without bilateral symmetry, the perceptual solution is always and only a triangle. The new effect found here, where symmetry strengthens the figure-ground salience of surfaces from occlusion cues, which form through visual spatial interpolation across fragments within a narrow temporal window of processing [44, 45], fully expresses the adaptive logic of visual preference for symmetry. Symmetry enables perception-based decision making, and survival relevant responses to symmetry or lack thereof can be found in animal species [37]. These highlight the wider biological significance of symmetry as a visual signal. The early Gestalt theories intuitively captured this fundamental importance in a large body of observations on phenomena of human perceptual organization. Their intuitions were astute, pointing towards functional aspects of symmetry which perception science has only just begun to quantify and predict. Funding: This research received no external funding. Acknowledgments: The experiments were materially supported by the CNRS. The contribution of Cyrielle Giger-Dechavanne, who helped recruiting subjects for the experiments, is gratefully acknowledged. Conflicts of Interest: The author declares no conflict of interest References 1. Wertheimer, M. Perceived Motion and Figural Organization, English trans. L. Spillmann, M. Wertheimer, K. W. Watkins, S. Lehar and V. Sarris MIT Press: Cambridge, 2012. 2. Metzger, W. Gesetze des Sehens, English trans. L. Spillmann Laws of Seeing MIT Press: Cambridge, USA 2009. 3. Deregowski, J. B. Symmetry, Gestalt and information theory. Quarterly Journal of Experimental Psychology 1991, 23, 381-385. 4. Gerbino, W.; Zhang, L. Visual orientation and symmetry detection under affine transformations. Bulletin of the Psychonomic Society 1991, 29, 480. 5. Machilsen, B., Pauwels, M.; Wagemans, J. The role of vertical mirror symmetry in visual shape perception. 6. Journal of Vision 2009, 9(11). Li, Y.; Sawada, T.; Shi, Y.; Steinman, R.M.; Pizlo, Z. (2013). Symmetry is the sine qua non of shape. In: S. Dickinson and Z. Pizlo (Eds.), Shape perception in human and computer vision. London: Springer 2013, pp.21-40. 7. Kanizsa, G. Organization in vision: Essays on Gestalt perception; Praeger: New York, USA, 1979. 8. Victor, JD; Conte, M. Illusory contour strength does not depend on the dynamics or relative phase of the 9. inducers. Vision Research 2000, 40, 3475 -- 3483. Spillmann, L.; Dresp, B. Phenomena of illusory form: Can we bridge the gap between levels of explanation? Perception 1995, 24, 1333 -- 1364. 10. Dresp, B. On "illusory" contours and their functional significance. Curr Psychol Cogn 1997, 16, 489 -- 518. 11. Shipley, T. F.; Kellman, P. J. Strength of visual interpolation depends on the ratio of physically specified- to-total edge length. Perception & Psychophysics 1992, 52, 97-106. 12. Dresp, B; Fischer, S. Asymmetrical contrast effects induced by luminance and color configurations. Perception & Psychophysics 2001, 63(7), 1262-1270. 13. Reid, R. C.; Shapley, R. M. (1988). Brightness induction by local contrast and the spatial dependence of assimilation. Vision Research 1988, 28, 115-132. 14. Gerrits, H. J. M.; Vendrik, A. J. Simultaneous contrast, filling-in process, and information processing in man's visual system. Experimental Brain Research 1970, 11, 411 -- 430. 15. Heinemann, E. G. Simultaneous brightness induction as a function of inducing and test-field luminance. Journal of Experimental Psychology 1955, 50, 89 -- 96. 16. von der Heydt, R. Figure -- ground organization and the emergence of proto-objects in the visual cortex. Front. Psychol. 2015, 6, 1695. Special issue on 'Perceptual Grouping-the State of the Art'. 17. Kellman, P.J.; Yin, C.; Shipley, T.F. A common mechanism for illusory and occluded object completion. Journal of Experimental Psychology: Human Perception & Performance 1998, 24(3), 859-869. 18. Kellman, P.J. (2002). Vision - occlusion, 'illusory' contours and filling-in. In Encyclopedia of Cognitive Science, Oxford, UK: Nature Publishing Group. 19. Yin, C.; Kellman, P.J.; Shipley, T.F. Surface completion complements boundary interpolation. Perception, Special Issue on surface appearance 1997, 26, 1459-1479. 20. Yin, C., Kellman, P.J.; Shipley, T.F. Surface integration influences depth discrimination. Vision Research 2000, 40(15), 1969-1978. 21. Kellman, P.J. From chaos to coherence: How the visual system recovers objects. Psychological Science Agenda of the American Psychological Association, 1997, 10(4), 8-9. 22. Dresp-Langley, B., Grossberg, S.; Reeves, A. Editorial: Perceptual Grouping -- The State of The Art. Front. Psychol. 2017, 8, 67. Special issue on 'Perceptual Grouping-the State of the Art'. 23. Shapley, R.; Gordon, J. Nonlinearity in the perception of form. Perception and Psychophysics 1985, 37, 84 -- 88. 24. Shapley, R.; Reid, R. C. Contrast and assimilation in the perception of brightness. Proceedings of the National Academy of Science 1985, 82, 5983 -- 5986. 25. Grossberg, S. 3D vision and figure-ground separation by visual cortex. Perception and Psychophysics 1994, 55, 48-120. 26. He, Z. J.; Ooi, T. L. Luminance contrast polarity: a constraint for occlusion and illusory surface perception. Investigative Ophthalmology and Visual Science 1998, 39 (Suppl.), 848. 27. Nakayama, K.; Shimojo, S.; Ramachandran, V. S. Transparency, relation to depth, subjective contours, luminance and neon color spreading. Perception 1990, 19, 497 -- 513. 28. Guibal, C. R. C.; Dresp, B. Interaction of color and geometric cues in depth perception: When does red mean near? Psychological Research 2004, 10: 167 -- 178. 29. Dresp-Langley, B.; Grossberg, S. Neural computation of surface border ownership and relative surface depth from ambiguous contrast inputs. Frontiers in Psychology 2016, 7, 1102. Special issue on 'Perceptual Grouping-the State of the Art'. 30. Hu, Q.; Victor, J.D. Two-Dimensional Hermite Filters Simplify the Description of High-Order Statistics of Natural Images. Symmetry 2016, 8, 98. 31. Bertamini, M; Griffin, L. Symmetry in Vision, MDPI: Basel, Switzerland, 2017, pp. 1-198. 32. Gillam, B. Redundant Symmetry Influences Perceptual Grouping (as Measured by Rotational Linkage). Symmetry 2017, 9, 67. 33. Mach, E. On Symmetry, In Popular Scientific Lectures, 1893; Lasalle: Open Court Publishing. 34. Wagemans, J. Characteristics and models of human symmetry detection. Trends in Cognitive Sciences, 1997 9, 346-352. 35. Grossberg, S. Cortical dynamics of figure-ground separation in response to 2D pictures, and 3D scenes: How V2 combines border ownership, stereoscopic cues, and Gestalt grouping rules. Front. Psychol. 2016, 6, 2054. Special issue on 'Perceptual Grouping-the State of the Art'. 36. Dresp-Langley, B. Affine Geometry, Visual Sensation, and Preference for Symmetry of Things in a Thing. Symmetry 2016 8(11), 127. 37. Giurfa, M.; Eichmann, B.; Menzl, R. Symmetry perception in an insect. Nature 1996, 382, 458 -- 461. 38. Bertamini, M. Sensitivity to reflection and translation is modulated by objectness. Perception 2010; 39, 27 -- 40. 39. Makin, A.J.D.; Rampone, G.; Wright, A.; Martinovic, J.; Bertamini, M. Visual symmetry in objects and gaps. Journal of Vision 2014, 14(3):12, 1 -- 12. 40. Baylis, G.C.; Driver, J. Perception of symmetry and repetition within and across visual shapes: Part-descriptions and object-based attention. Visual Cognition 2001, 8(2), 163 -- 196. 41. Michaux, A; Kumar, V.; Jayadevan, V.; Delp, E., Pizlo, Z. Binocular 3D Object Recovery Using a 42. Symmetry Prior. Symmetry 2017, 9, 64. Jayadevan V., Sawada, T; Delp, E; Pizlo, Z. Perception of 3D Symmetrical and Nearly Symmetrical Shapes. Symmetry 2018, 10, 344. 43. Poirier, F.J.A.M.; Wilson, H.R. A biologically plausible model of human shape symmetry perception. Journal of Vision, 2010,10(1):9, 1 -- 16. 44. Erlikhman, G.; Kellman, P.J. From flashes to edges to objects: Recovery of local edge fragments initiates spatiotemporal boundary formation. Frontiers in Psychology 2016, 7, 910. Special issue on 'Perceptual Grouping-the State of the Art'. 45. Erlikhman, G.; Kellman, P.J. Modeling spatiotemporal boundary formation. Vision Research 2015, 126, 131-142.
1606.02372
1
1606
2016-06-08T02:01:20
Entrainment of traveling waves to rhythmic motor acts
[ "q-bio.NC" ]
We hypothesized that rhythmic motor acts entrain neural oscillations in speech production brain regions. We tested this hypothesis in an experiment where a subject produced consonant-vowel (CV) syllables in a rhythmic fashion, while we performed ECoG recordings. Over the ventral sensorimotor cortex (vSMC) we detected significant concentration of phase across trials at the specific frequency of speech production. We also observed amplitude modulations. In addition we found significant coupling between the phase of brain oscillations at the frequency of speech production and their amplitude in the high-gamma range (i.e., phase-amplitude coupling, PAC). Furthermore, we saw that brain oscillations at the frequency of speech production organized as traveling waves (TWs), synchronized to the rhythm of speech production. It has been hypothesized that PAC is a mechanism to allow low-frequency oscillations to synchronize with high-frequency neural activity so that spiking occurs at behaviorally relevant times. If this hypothesis is true, when PAC coexists with TWs, we expect a specific organization of PAC curves. We observed this organization experimentally and verified that the peaks of high-gamma oscillations, and therefore spiking, occur at the same times across electrodes. Importantly, we observed that these spiking times were synchronized with the rhythm of speech production. To our knowledge, this is the first report of motor actions organizing (a) the phase coherence of low-frequency brain oscillations, (b) the coupling between the phase of these oscillations and the amplitude of high-frequency oscillations, and (c) TW. It is also the first demonstration that TWs induce an organization of PAC so that spiking across spatial locations is synchronized to behaviorally relevant times.
q-bio.NC
q-bio
Entrainment of traveling waves to rhythmic motor acts Joaqu´ın Rapela∗ Swartz Center for Computation Neuroscience University of California San Diego September 25, 2018 Abstract A remarkable early observation on brain dynamics is that when hu- mans are exposed to rhythmic stimulation their brain oscillations entrain to the rhythm of the stimuli (Adrian and Matthews, 1934). Currently it is not know whether rhythmic motor acts can entrain neural oscilla- tions. We investigated this possibility in an experiment where a subject produced consonant-vowel (CV) syllables in a rhythmic fashion, while we performed ECoG recordings with a dense grid covering most speech pro- cessing brain regions across the left hemisphere. Most strongly over the ventral sensorimotor cortex (vSMC), a cortical region that controls the vocal articulators, we detected significant concentration of phase across trials at the specific frequency of speech production. We also observed am- plitude modulations. In addition we found significant coupling between the phase of brain oscillations at the frequency of speech production and their amplitude in the high-gamma range (i.e., phase-amplitude coupling, PAC). Furthermore, we saw that brain oscillations at the frequency of speech production organized as traveling waves (TWs), synchronized to the rhythm of speech production. It has been hypothesized that PAC is a mechanism to allow low-frequency oscillations to synchronize with high- frequency neural activity so that spiking occurs at behaviorally relevant times. If this hypothesis is true, when PAC coexists with TWs, we expect a specific organization of PAC curves. We observed this organization ex- perimentally and verified that the peaks of high-gamma oscillations, and therefore spiking, occur at the same times across electrodes. Importantly, we observed that these spiking times were synchronized with the rhythm of speech production. To our knowledge, this is the first report of motor actions organizing (a) the phase coherence of low-frequency brain oscil- lations, (b) the coupling between the phase of these oscillations and the amplitude of high-frequency oscillations, and (c) TWs. It is also the first demonstration that TWs induce an organization of PAC so that spiking across spatial locations is synchronized to behaviorally relevant times. ∗[email protected] Glossary CV consonant vowel. ECoG electrocorticography. A brain recording modality that measures poten- tials directly from the cortical surface. EEG electroencephalography. A brain recording modality that measures elec- trical potentials from the scalp.. ERP event-related potential. Mean of voltages recorded at a single electrode and aligned with respect to an event of interest (e.g., the transition time between a consonant and a vowel).. ERSP event-related spectral perturbation. An average measure of evoked power across trials (Section A.1.3). ITC inter-trial coherence. A measure of phase alignment between trials (Sec- tion A.1.2). MEG magnetoencephalography. A brain recording modality that measures magnetic fields from the scalp.. MI modulation index quantifying the strength of phase-amplitude coupling (Sections 2.3 and A.1.4). PAC phase-amplitude coupling. Coupling between the phase of low-frequency oscillations and the amplitude of high-frequency ones (Sections 2.3 and A.1.4). TW traveling wave (Section 2.4). vSMC ventral sensorimotor cortex. A brain region that controls the vocal articulators. Contents 1 Introduction 2 Results 2.1 Phase Alignment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.2 Amplitude Modulations 2.3 Phase-Amplitude Coupling (PAC) . . . . . . . . . . . . . . . . . 2.4 Traveling Waves (TWs) . . . . . . . . . . . . . . . . . . . . . . . 2.5 Organization of PAC in the Presence of TWs . . . . . . . . . . . 3 Discussion 4 Acknowledgments A Supplementary Information A.1 Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . A.1.1 Circular statistics concepts A.1.2 ITC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . A.1.3 ERSP . . . . . . . . . . . . . . . . . . . . . . . . . . . . . A.1.4 Phase-amplitude coupling curve and modulation index . . A.2 Behavioral Data . . . . . . . . . . . . . . . . . . . . . . . . . . . A.3 ITC across the vSMC . . . . . . . . . . . . . . . . . . . . . . . . A.4 ERSPs across the vSMC . . . . . . . . . . . . . . . . . . . . . . . A.5 ERPs across the vSMC . . . . . . . . . . . . . . . . . . . . . . . A.6 Electrodes with strongest PAC . . . . . . . . . . . . . . . . . . . 1 2 2 4 5 7 9 10 13 2 2 2 2 2 3 3 3 16 34 34 1 Introduction External visual (Regan, 1966) and auditory (Galambos et al., 1981) rhythmic stimuli entrain brain oscillations (i.e., drag brain oscillations to follow the rhythm of the stimuli), and this entrainment is modulated by attention so that the oc- currence of attended stimuli coincides with the phase of brain oscillations of maximal excitability (Lakatos et al., 2005, 2008, 2013; O'Connell et al., 2011; Besle et al., 2011; Gomez-Ramirez et al., 2011; Zion Golumbic et al., 2013; Cravo et al., 2013; Mathewson et al., 2010; Spaak et al., 2014; Gray et al., 2015). Although speech is only quasi rhythmic (Cummins, 2012) an increasing number of studies is showing that neural oscillations can entrain to speech sound (e.g., Zion Golumbic et al., 2013; Gross et al., 2013; Doelling et al., 2014; Millman et al., 2015; Park et al., 2015). For recent reviews on the entrainment of neural oscillations to speech sound see Peelle and Davis (2012); Zion Golumbic et al. (2012); Ding and Simon (2014). What is currently not know if whether speech production can entrain neural oscillations. Here we present evidence showing that a speaker rhythmi- cally producing consonant-vowel (CV) syllables at different frequencies entrains brain oscillations in the ventral sensorimotor cortex (vSMC) at corresponding different frequencies. It has been suggested that a role of entrained low-frequency oscillations is to modulate the excitability of neurons, in such a way that periods of higher ex- citability correspond to events of interest in sensory streams (Schroeder and Lakatos, 2009). This hypothesis has been supported by reports of phase-amplitude cou- pling (PAC), where the phase of low-frequency oscillations modulates the ampli- tude of higher-frequency ones (e.g., Canolty et al., 2006). In speech perception it has been proposed that theta rhythm (3-8 Hz) is related to the encoding of slower syllabic information in the speech signal, the gamma rhythm (>30 Hz) is involved in the linguistic coding of phonemic details, and the PAC between the theta and gamma rhythms could modulate the excitability of neurons to devote more processing power to the informative parts of syllabic sound pat- terns (Giraud and Poeppel, 2012; Hyafil et al., 2015). However, PAC has not yet been reported in speech-production brain regions. We show that over the vSMC the low-frequency oscillations entrained to the rhythm of speech pro- duction are strongly coupled with high-gamma oscillations. Importantly, we demonstrate that the phase at which the entrained oscillation couples with the maximum amplitude of the high-gamma oscillations changes orderly from ven- tral to dorsal electrodes over the vSMC. Traveling waves (TWs) have been reported in animal studies (e.g., Rubino et al., 2006), human EEG (e.g., Patten et al., 2012), and human ECoG (e.g., Bahramisharif et al., 2013). However, these waves have not been observed in speech production brain regions in humans. The aforementioned orderly change of PAC across the vSMC suggested the existence of TWs. We looked for them and we found them. As mentioned above, brain activity at single recording sites can be entrained to external stimulation but, to our knowledge, it is unknown whether distributed spatio-temporal activity in the form of TWs can also be entrained. We demon- strate that TWs over the vSMC are entrained to the rhythm of speech produc- 1 tion. PAC has been studied in the absence of TWs. When these waves are present, in order for cells to fire at the same behaviorally-relevant time, PAC curves mea- sured across electrodes should be different from each other and systematically organized. If these curves were identical to each other, cells would fire in the or- der given by the propagating TW, and not at the same time. Below we propose an organization of PAC so that when TWs are present cells fire at the same time, we show that this organization is present in our recordings, and provide evidence indicating that cells along the direction of propagation of a planar TW fire at the same behaviorally relevant time. 2 Results 2.1 Phase Alignment We epoched the raw ECoG waveforms around the time of CV transitions (rep- resented by the vertical line at time zero in all figures). We analyzed separately recordings from different sessions, and present the analysis from recording ses- sions EC2 B105 and EC2 B89, the sessions with the largest (1.62 sec) and short- est (0.97 sec) median inter-syllable separation times, respectively (Figure A.1). Figure 1a and 1b show the inter-trial coherence (ITC, Section A.1.2), a measure of phase alignment in a group of trials, calculated from recordings in sessions EC2 B105 and EC2 B89, respectively, for electrode 154 in the vSMC (Figure 2). In agreement with our hypothesis that the rhythm of speech production entrains low-frequency brain oscillations, in the ITC computed from data from experi- mental session EC2 B105 with a median inter-syllable production frequency of 0.62 Hz (black horizontal line in Figure 1a), we observe a peak of ITC around this frequency (red pixels in Figure 1a). Also, we see a peak of ITC around a frequency of 1.03 Hz in Figure 1b, obtained from an experimental session where the median inter-syllable separation time was 1.03 Hz. These figures il- lustrate that low-frequency brain oscillations entrain to the rhythm of speech production. For each electrode, we computed the maximum ITC between 0.5 seconds before and 0.6 seconds after the CV transition at the frequency of speech pro- duction. Figure 2 highlights the 50 electrodes with largest ITCs. We sorted the electrodes by their maximum ITC value, and colored them according to their rank in this sorting. The ten electrodes with largest ITC are colored in red and are located in the center of the ventral sensoritmotor cortex. As the distance of an electrode to the center of the vSMC increases, its maximum ITC decreases. Section A.3 more clearly illustrates the specificity of phase coherence to the vSMC by showing ITC plots from the ventral to the dorsal edge of the grid across the vSMC. 2 Figure 1: ITCs for electrode 154 computed from recordings in experimental session EC2 B105 with an inter-syllable production frequency of 0.62 Hz (a), and from experimental session EC2 B89 with an inter-syllable production frequency of 1.03 Hz (b). Note that the peaks of the ITC at the lowest frequencies (red points) align with the median frequency of speech production (black horizontal line) both when the subject speaks slower (a) and faster (b). 3 Figure 2: Strongest entrainment occurs over the vSMC. The 50 electrodes with largest ITC (between 0.5 seconds before and 0.6 seconds after the CV transition, at the frequency of CV production, in the experimental session EC2 B105) are highlighted in color. 2.2 Amplitude Modulations The Event-Related Spectral Perturbation (ERSP, Section A.1.3) is a measure of power modulations around an event of interest. Figure 3a shows ERSPs around the transition time between consonants and vowels (black vertical line in Figure 3). In electrode 133 over auditory cortex (Figure 3a) we see a significant power increase (red blob) before the transition between consonants and vowels in the beta range. Differently, in electrode 135 over the vSMC (Figure 3b) we observe a significant power increase after the transition between consonants and vowels and in the high-gamma range. Figures A.18-A.30 show ERSPs between the ventral electrode 129 and the dorsal electrode 141 in experimental session EC2 B105. Power in the beta range increases from electrodes 129 to 132 (Figures A.18-A.21) in the ventral tempo- ral cortex, and peaks over electrodes 133 and 134 around the auditory cortex (Figures A.22 and A.23), prior to consonant-vowel transitions. A sharp transi- tion is seen in electrode 135 over the ventral sensorimotor cortex (Figure A.24), with strong modulations in the high-gamma range (70-200 Hz), after consonant- vowel transitions, and almost no modulation in the beta range. This pattern of modulations is strongest over electrode 136 (Figure A.25), and fades along electrodes 137-141 (Figures A.26-A.30) toward the dorsal end of the vSMC. 4 (a) ERSP in Auditory Cortex (b) ERSP in the vSMC Figure 3: ERSPs for electrode 133 in the auditory cortex (a) and for electrode 135 in the vSMC (b). In the auditory cortex (a) we observe a peak of amplitude (red blob) in the beta range before the CV transition (vertical black line), while in the vSMC (b) a peak of amplitude occurs in the high-gamma range after the CV transition. 2.3 Phase-Amplitude Coupling (PAC) For experimental session EC2 B105, we studied the coupling between phase at the entrainment frequency of 0.62 Hz (0.01 Hz, wavelet full width at half maxi- mum) and amplitude at 100 Hz (2.06 Hz, wavelet full width at half maximum). For each electrode we calculated the PAC curve (Section A.1.4), and quantified the strength of the coupling using the modulation index (MI, Section A.1.4). Figure 4 shows PAC curves measured at electrodes 135-138 (for visualization purposes we show two phase cycles). The ECoG power at 100 Hz is well mod- ulated by the phase at the entrainment frequency, with MIs 4.84e-3, 8.22e-3, 1.05e-3, and 2.82e-3 at electrodes 135-138, respectively. For electrode 136 (green trace) larger ECoG amplitudes in the high-gamma range (100 Hz) tend to oc- cur at the valley (phase near π) of the low-frequency oscillation (0.62 Hz). As we move from the more ventral electrode 135 to the more dorsal electrode 138, the phase of the low-frequency oscillation at which the amplitude of the high- gamma oscillations is maximal shifts gradually from π (pink trace) to 0 (violet trace). We discuss this organization in Section 2.5. Strongest PAC between phase at the entrainment frequency and amplitude at 100 Hz was observed over the vSMC and surrounding regions. Figure A.44 highlights the 50 electrodes with largest MI. For electrode 154 in the vSMC, we computed the MI for all pairs of frequency for phase and frequency for amplitude, with the frequency for phase between 5 ) V ( z H 0 0 1 t a e d u t i l p m A n a e M 3 2 Electrode 138 137 136 135 2p 0 Phase (radians) 3p Figure 4: Phase-amplitude coupling curves at electrodes over the vSMC. Phase and amplitude were measured at the entrainment frequency (0.62 Hz) and at 100 Hz, respectively. All displayed electrodes showed a strong modulation of the high-gamma amplitude by the entrained low-frequency phase. Note that as we move from more ventral to more dorsal electrodes (i.e., from electrode 135 to electrode 138) the peak of the PAC curves occurs at earlier phases. The significance of this organization is discussed in Section 2.5. 6 - p p m 0.1 and 20 Hz, and the frequency for amplitude between 5 and 200 Hz. We only observed large MIs for frequencies for phase between 0.1 and 2 Hz. Figure 5a plots the obtained MIs. We found two clusters of the frequency for amplitude with large MIs. The first cluster was on the high-gamma range (60 to 150 Hz) and the second cluster on the beta range (20 to 40 Hz). An asterisk in Figure 5 marks a MI peak, the number next to the asterisk is the peak MI, and the first and second numbers in the parenthesis are the frequency for phase and for amplitude, respectively, of the peak MI. Interestingly, for the cluster on the high- gamma range, the maximum MI occurred exactly at the bin of the entrainment frequency (0.62 Hz for the frequency for phase). Differently, for the cluster in beta range the maximum MI occurred below the entrainment frequency (0.5 Hz for the frequency for phase). This figure suggests that not only low-frequency oscillations become entrained to the rhythm of speech production, but also PAC is optimized at this rhythm. For electrode 132 in the auditory cortex, Figure 5b, the MIs were overall much weaker, and both the clusters in the high-gamma and beta ranges showed a peak MI at frequencies for phase below the entrainment frequency. This suggest that the maximization of PAC at the entrainment frequency is specific to the vSMC. 2.4 Traveling Waves (TWs) We filtered the ECoG recordings in experimental session EC2 B105 (Figure A.1) with a second-order bandpass Butterworth filter around the median syllable- production syllable of 0.62 Hz (low- and high-frequency cutoffs 0.4 Hz and 0.8 Hz, respectively). Figure 6 shows the filtered voltages from electrodes 135- 141 at the center of the vSMC in a representative time interval (390 to 400 sec- onds after the start of the experiment). We see that, as we move from the ventral electrode 135 to the dorsal electrode 141, the voltage traces are orderly shifted to later times. The vertical black lines indicate the transition time between a consonant and a vowel in the production of consonant-vowel syllables. Note that the peaks of the voltage traces tend to occur around these consonant-vowel transition times. Thus, this figure illustrates a wave of activity traveling from the ventral to the dorsal edge of the ventral sensorimotor cortex and entrained to the rhythm of speech production. A movie showing the cosine of the phase of the bandpassed filtered voltages across the whole grid simultaneously with the speech of the subject can be found at https://youtu.be/dXrzj2eEuVY. The bottom/top/left/right pixels in the movie correspond to electrodes along the ventral/dorsal/frontal/caudal sides of the grid. We see mostly planar traveling waves moving in different directions, but occasionally we also observe rotating traveling waves. At most times when the subject produces a CV syllable we see a white blob over the vSMC, indicating that filtered voltages approach their peak values. This movie demonstrates, for the first time, the existence of traveling waves across left speech processing brain areas synchronized to the rhythm of speech production. 7 (a) (b) 200 150 e d u t i l p m A 100 r o f y c n e u q e r F * 0.0029 (0.70, 87.00) * 0.0024 (0.50, 27.00) 50 0 200 150 100 50 0 MI 0.003 0.002 0.001 0.000 * 0.0004 (0.50, 63.00) * 0.0010 (0.40, 16.00) 0.5 1.0 1.5 2.0 Frequency for Phase 0.5 1.0 1.5 Frequency for Phase 2.0 Figure 5: Modulation indices for electrode 154 over the vSMC (a) and for electrode 132 over the auditory cortex (b) at multiple frequencies for phase and amplitude. Over the vSMC (a) the peak MI for amplitudes in the high-gamma range occurs at the entrained frequency for phase (0.62 Hz), while the peak MI for amplitudes in the beta range occurs at lower frequencies. This suggests that over the vSMC not only phase coherence, but also phase-amplitude coupling, is matched to the speech production frequency. Over auditory cortex (b) MIs are much smaller than over the vSMC and the peak MIs occurred at frequencies for phase below the entrainment frequency. This suggests that PAC is largest over speech-production brain regions and that the maximization of PAC at the entrainment frequency is specific to the vSMC. 8 ) V ( e g a t l o V 20 10 0 −10 −20 −30 Electrode 138 137 136 135 390.0 392.5 395.0 Time (sec) 397.5 400.0 Figure 6: Filtered voltage waveforms along the ventral-dorsal axis of the vSMC. Filtering was performed between 0.4 and 0.8 Hz, around the median frequency of CV syllable production of 0.62 Hz. Vertical black lines indicate CV transition times. As we move from the ventral electrode 135 in red to the dorsal electrode 141 in magenta, voltage waveforms are orderly shifted to right, indicating the existence of a traveling wave. Also, the peak of the waves occurs around the time of CV transitions, suggesting that these traveling waves are entrained to the rhythm of speech production. 2.5 Organization of PAC in the Presence of TWs If we wanted to design a brain so that all neurons recorded by the four electrodes in Figure 4 fired at the same behaviorally relevant time (e.g., the transition between the consonant-vowel syllable at time 391.2 seconds), how would be choose the PAC curves for these electrodes? Since the phase at electrode 135 (red trace in Figure 4) at time 391.2 is close to π/4 we would use a PAC curve for this electrode with maximal high-gamma amplitude at this phase. Then, because high-gamma amplitude is related to neural firing (Ray et al., 2008), cells around electrode 135 would then tend to spike at the time of the consonant vowel transition. Next, due to the traveling wave, the phase at electrode 136 at time 391.2 is earlier than that at electrode 135, thus we would use a PAC curve with a peak at this earlier phase for electrode 136. Similarly, we would use a PAC curve with an earlier peak at electrode 137 than that at electrode 136. That is, as we move along the direction of propagation of a traveling wave, we would use PAC curves with earlier and earlier peaks. And this is what we observed experimentally in Figure 4. To validate that the previous organization of PAC curves allows cells to fire at the same behaviorally relevant time in the presence of traveling waves, Fig- ure 7 plots mean high-gamma amplitude as a function of time for electrodes 135-138 (for each time and each electrode, we extracted the phase at 0.62 Hz, then the mean amplitude at 100 Hz for this phase from the PAC curve in Fig- ure 4, and finally we plotted the extracted mean amplitudes as a function of the corresponding times). As expected, we found that high-gamma amplitude, and 9 m ) V (m z H 0 0 1 t a e d u t i l p m A n a e M 3 2 Electrode 138 137 136 135 390.0 392.5 395.0 Time (sec) 397.5 400.0 Figure 7: Mean high-gamma amplitude as a function of time at electrodes over the ventral sensorimotor cortex. Each time point at each electrode corresponds to a given phase of the oscillation at the speech-production frequency of 0.62 Hz in Figure 6, and from Figure 4 each phase corresponds to a given mean amplitude at 100 Hz. Thus, each time point of each electrode corresponds to a mean phase at 100 Hz. Colored traces show this correspondence across different electrodes. We observe that mean high-gamma amplitude, and therefore spiking, is aligned across electrodes over the ventral sensorimotor cortex and synchronized to the rhythm of speech production. The curves are not smooth because we used coarse bins in the calculation of the PAC curves in Figure 4. therefore cell spiking, peaks at approximately similar times across electrodes. In addition, we observed that the high-gamma amplitude peaks occur shortly after the transition time between consonant and vowels, which shows that the traveling waves are therefore entrained to the rhythm of speech production. 3 Discussion We have shown that rhythmic speech production entrains low-frequency brain oscillations (Figure 1) and that this entrainment is stronger over the vSMC (Figure 2), that rhythmic speech production modulates the power of beta os- cillations over the auditory cortex and that of high-gamma oscillations over the vSMC (Figure 3), that the phase of the entrained low-frequency oscillations couples with the amplitude of high-gamma oscillations (Figure 4) and that this coupling is largest over the vSMC (Figure A.44), that over the vSMC -- but not over the auditory cortex -- this coupling is maximal between phase at the entrained frequency and amplitude in the high-gamma range (Figure 5), and that the entrained oscillations organize as TWs synchronized to the produced speech (Figure 6 and movie at https://youtu.be/dXrzj2eEuVY). Finally, we observed that the peaks of PAC curves occur at earlier phases along the direc- tion of propagation of the TWs and presented evidence (Figure 7) indicating that this organization of PAC curves enables neurons along the direction of 10 propagation of the TW to fire at the same behaviorally relevant time. These initial findings open several interesting directions for future investi- gations. To mention a few: (1) How do entrainment, PAC, and TWs relate to properties of the produced CV syllables? For example, does PAC occur earlier for fricatives than for nasals? (2) How do these effects (i.e., entrain- ment, PAC, and TWs) correspond to the specific articulators recruited in the production of a CV syllable? For example, do planar TWs correspond to the recruitment of the jaw but rotating TWs to the recruitment of the larynx? (3) PAC has been reported in speech perception (e.g., Giraud and Poeppel, 2012; Hyafil et al., 2015), but not in speech production tasks. We claimed that the PAC we observed is related to speech production, but we have not eliminated the possibility that it may be related to the perception of the speaker's own speech. The detection of PAC in new experiments eliminating speech feedback would confirm that PAC can be related to the production of speech. (4) The observed effects appeared in the strongly rhythmic production of CV syllables. Would similar effects be observed in colloquial speech? How do prosodic fea- tures of speech, like intonation or stress, affect the observed effects? (5) We have characterized phase coherence and PAC at single electrodes, but probably phase coherence and PAC across different electrodes play an important role in the synchronization and information transfer across different speech perception and production brain regions, as observed in non-speech-processing tasks (e.g., Voytek et al., 2015). It is possible to explain some of our results in terms of event-related poten- tials (ERPs, the mean recorded potential around an event of interest). Previous studies have suggested that ERPs are generated by the alignment of phases of ongoing oscillations (Makeig et al., 2002), but this suggestion is still controver- sial (Sauseng et al., 2007). Because the production of CV syllables occurred at a low-frequency rhythm, ERPs should be generated at this low-frequency and we should observe alignment of phases at this low frequency. Thus, this ERP argument explains Figure 1. Also, since ERPs should be stronger over the vSMC, we should expect more phase alignment over this region, explaining Figure 2. One could argue that the production of speech is related to spiking, and therefore to high-gamma activity, over the vSMC. Thus, the low frequency phase at which oscillations reset for the generation of ERPs should be related to high-gamma activity. This argument addresses Figures 4 and 5. Section A.5 shows ERPs computed along the ventral-dorsal axis of the recording grid. In agreement with Figure 2, we see strongest ERPs at elec- trodes 136-140 over the vSMC. These ERPs begin around 150 msec before the CV transition, peak around 100 ms after this transition, and show a negative peak around 400 ms after the CV transition. Also, in agreement with Figure 4 we see that the peak of the ERPs are shifted to later times as we move from electrode 136 to electrode 140, also suggesting the existence of TWs. Thus, the spectral and ERP analysis are consistent with each other. An advantage of the former is that it decomposes the averaged effects seen in ERPs in different frequency bands. Since several frequency-specific effects on electro- physiological recordings are known, spectral analysis allow richer descriptions 11 of electrophysiological recordings than ERP ones. For instance by decomposing averaged effects in different frequency bands, our spectral analysis showed that neural spiking, as reflected by high-gamma amplitude, was coupled to relevant times in the speech production rhythm, as indicated by the coupled phase at the speech-production frequency (Figure 4). Also, our PAC analysis is more quantitative than the previous argument linking ERPs and high-gamma ampli- tude. This analysis revealed a new organization of PAC in the presence of TWs, providing strong support to a previous hypothesis stating that, when stimuli or behavior occurs in a rhythmic fashion, PAC is a mechanism that allows neurons to spike at behavioral relevant times (Canolty and Knight, 2010). In summary, the spectral methods used above to characterize rhythmic speech production are consistent with, but superior than, ERP methods for the characterization of rhythmic behaviors such as speech. The effects reported above reflect brain mechanisms beyond the control of vo- cal articulators. For example, the traveling wave movie (https://youtu.be/dXrzj2eEuVY) shows waves moving across multiple speech processing brain regions. At sev- eral times we see waves moving from electrodes over the auditory cortex to electrodes over the vSMC, suggesting that these waves could be a mecha- nism for integrating information across the brain, as suggested by others (e.g., Nunez and Srinivasan, 2006; Sato et al., 2012). As speaking must be compatible with the performance limits of motor sys- tems, several theories have linked functional properties of motor systems to speech capability. The Frame/Content theory (MacNeilage and Davis, 2000, 2001) proposes that the features of elementary production units, i.e., sylla- bles, are determined by mechanical properties of the speech apparatus, e.g., natural oscillatory rhythms (Giraud et al., 2007). Previous investigations (e.g., Morillon et al., 2010) have hypothesized that motor areas express low-frequency oscillatory activity characteristic of jaw movements (4 Hz) and high-frequency activity corresponding to movements of the tongue (e.g., trill at 35-40 Hz). How- ever, to our knowledge this is the first study providing physiological evidence for the relation of low- and high-frequency oscillations, as well as their interactions, to speech production. Natural speech production is substantially more complex than the produc- tion of CV syllables, which suggests that the findings reported in this article may not be relevant to natural speech. We argue on the contrary. CV sylla- bles are the building blocks of natural speech, such as sinusoidal grating are the building blocks of natural images. While producing CV syllables subjects spontaneously enter a strongly rhythmic behavioral state. Natural speech is strongly rhythmic and the production of CV syllables could be ideal to isolate rhythmic aspects of this behavior. By characterizing responses of visual cells to sinusoidal gratings, it has been possible to understand key neural mechanisms of visual processing, like contrast grain control (Ohzawa et al., 1985), that are difficult to observe in responses of visual neurons to complex natural images. Similarly, investigating the production of CV syllables may be instrumental in discovering key rhythmic mechanisms of speech production that are obscured in more complex speech-production tasks. 12 The specific attributes of mesoscopic cortical activity related to speech pro- duction are not well characterized. In speech perception, Luo and Poeppel (2007) have shown that phase activity (over the human auditory cortex in the theta range (4-8 Hz) recorded with MEG) reliably tracks and discriminates spoken sentences and that this discrimination is correlated with speech intelli- gibility. Here we extended these results and showed that phase activity (in the form of phase alignment, PAC, and TWs) is also a relevant mesoscopic attribute to characterize the neurophysiology of speech production. Having detected phase alignment, PAC and TWs in ECoG recordings, one could use forward models to calculate how these effects propagate to scalp- recorded potentials in the EEG or magnetic fields in the MEG. Next, one could invert these forward models to infer the existence and properties of these oscil- latory effects from EEG or MEG recordings (Friston et al., 2007). If successful this approach would allow to detect phase alignment, PAC, and TWs from non- invasive recordings, and provide novel measures to non-invasively characterize speech production in health and disease. 4 Acknowledgments We thank Dr. Edward Chang and Dr. Kristofer Bouchard for sharing the ECoG recordings with us and for comments on a preliminary version of this manuscript, and Dr. John Iversen for suggesting us the relation between PAC and TWs. 13 References E.D. Adrian and B.H.C. Matthews. The berger rhythm : potential changes from the occipital lobes in man. Brain, 57(4):355 -- 385, 1934. Ali Bahramisharif, Marcel AJ van Gerven, Erik J Aarnoutse, Manuel R Mercier, Theodore H Schwartz, John J Foxe, Nick F Ramsey, and Ole Jensen. Propagating neocortical gamma bursts are coordinated by traveling alpha waves. The Journal of Neuroscience, 33(48):18849 -- 18854, 2013. J. Besle, C.A. Schevon, A.D. Mehta, P. Lakatos, R.R. Goodman, G.M. McKhann, R.G. Emerson, and C.E. Schroeder. Tuning of the human neocortex to the temporal dynamics of attended events. The Journal of Neuroscience, 31(9):3176 -- 3185, 2011. R.T. Canolty and R.T. Knight. The functional role of cross-frequency coupling. Trends in cognitive sciences, 14(11):506 -- 515, 2010. R.T. Canolty, E. Edwards, S.S. Dalal, M. Soltani, S.S. Nagarajan, H.E. Kirsch, M.S. Berger, N.M. Barbaro, and R.T. Knight. High gamma power is phase-locked to theta oscillations in human neocortex. science, 313(5793):1626 -- 1628, 2006. A.M. Cravo, G. Rohenkohl, V. Wyart, and A.C. Nobre. Temporal expectation en- hances contrast sensitivity by phase entrainment of low-frequency oscillations in visual cortex. Journal of Neuroscience, 33(9):4002 -- 4010, 2013. F. Cummins. Oscillators and syllables: a cautionary note. Frontiers in psychology, 3: 364, 2012. A. Delorme and S. Makeig. EEGLAB: an open source toolbox for analysis of single-trial EEG dynamics including independent component analsys. Journal of Neuroscience Methods, 134(1):9 -- 21, 2004. N. Ding and J.Z. Simon. Cortical entrainment to continuous speech: functional roles and interpretations. Front. Hum. Neurosci, 8(311):10 -- 3389, 2014. Keith B Doelling, Luc H Arnal, Oded Ghitza, and David Poeppel. Acoustic land- marks drive delta -- theta oscillations to enable speech comprehension by facilitating perceptual parsing. Neuroimage, 85:761 -- 768, 2014. K.J. Friston, J.T. Ashburner, S.J. Kiebel, T.E. Nichols, and W.D. Penny, editors. Statistical Parametric Mapping. Academic Press, London, 2007. Robert Galambos, Scott Makeig, and Peter J Talmachoff. A 40-hz auditory potential recorded from the human scalp. Proceedings of the National Academy of Sciences, 78(4):2643 -- 2647, 1981. A.L. Giraud, A. Kleinschmidt, D. Poeppel, T.E. Lund, R.S.J. Frackowiak, and H. Laufs. Endogenous cortical rhythms determine cerebral specialization for speech perception and production. Neuron, 56(6):1127 -- 1134, 2007. Anne-Lise Giraud and David Poeppel. Cortical oscillations and speech processing: emerging computational principles and operations. Nature neuroscience, 15(4):511 -- 517, 2012. 14 M. Gomez-Ramirez, S.P. Kelly, S. Molholm, P Sehatpour, T.H. Schwartz, and J.J. Foxe. Oscillatory sensory selection mechanism during intersensory attention to rhythmic auditory and visual inputs: a human electrocortigraphic investigation. The Journal of Neuroscience, 31(50):18556 -- 18567, 2011. Michael J Gray, Hans-Peter Frey, Tommy J Wilson, and John J Foxe. Oscillatory recruitment of bilateral visual cortex during spatial attention to competing rhythmic inputs. The Journal of Neuroscience, 35(14):5489 -- 5503, 2015. J. Gross, N. Hoogenboom, G. Thut, P. Schyns, S. Panzeri, P. Belin, and S. Garrod. Speech rhythms and multiplexed oscillatory sensory coding in the human brain. PLoS Biology, 11(12), 2013. doi: 10.1371/journal.pbio.1001752. A. Hyafil, L. Fontolan, C. Kabdebon, B. Gutkin, and A.L. Giraud. Speech encoding by coupled cortical theta and gamma oscillations. Elife, 4:e06213, 2015. P. Lakatos, A.S. Shah, K.H. Knuth, I. Ulbert, G. Kamos, and C.E. Schroeder. An oscillatory hierarchy controlling neural excitability and stimulus processing in the auditory cortex. J. Neurophysiology, 94(3):1904 -- 1911, 2005. P. Lakatos, G. Karmos, A.D. Mehta, I. Ulbert, and C.E. Schroeder. Entrainment of neuronal oscillations as a mechanism of attentional selection. Science, 320:110 -- 113, 2008. P. Lakatos, G. Musacchia, M.N. O'connel, A.Y. Falchier, D.C. Javitt, and C.E. Schroeder. The spectrotemporal filter mechanism of auditory selective attention. Neuron, 77(4):750 -- 761, 2013. Huan Luo and David Poeppel. Phase patterns of neuronal responses reliably discrim- inate speech in human auditory cortex. Neuron, 54(6):1001 -- 1010, 2007. P.F. MacNeilage and B.L. Davis. On the origin of internal structure of word forms. Science, 288(5465):527 -- 531, 2000. P.F. MacNeilage and B.L. Davis. Motor mechanisms in speech ontogeny: phylogenetic, neurobiological and linguistic implications. Current opinion in neurobiology, 11(6): 696 -- 700, 2001. S. Makeig, M. Westerfield, T.P. Jung, S. Enghoff, J. Townsend, E. Couchesne, and T.J. Sejnowski. Dynamic brain sources of visual evoked responses. Science, 295 (5555):690 -- 694, 2002. K.V. Mardia. Statistics of directional data. Acadmic Press, New York, NY, 1972. K.E. Mathewson, M. Fabiani, G. Gratton, D.M. Beck, and A. Lleras. Rescuing stimuli from invisibility: Inducing a momentary release from visual masking with pre-target entrainment. Cognition, 115:186 -- 191, 2010. R.E. Millman, S.R. Johnson, and G. Prendergast. The role of phase-locking to the temporal envelope of speech in auditory perception and speech intelligibility. Journal of cognitive neuroscience, 2015. 15 B. Morillon, K. Lehongre, R.S.J. Frackowiak, A. Ducorps, A. Kleinschmidt, D. Poep- pel, and A.L. Giraud. Neurophysiological origin of human brain asymmetry for speech and language. Proceedings of the National Academy of Sciences, 107(43): 18688 -- 18693, 2010. P.L. Nunez and R. Srinivasan. A theoretical basis for standing and traveling brain waves measured with human eeg with implications for an integrated consciousness. Clinical Neurophysiology, 117(11):2424 -- 2435, 2006. M.N. O'Connell, A. Falchier, T. McGinnis, C.E. Schroeder, and P. Lakatos. Dual mechanism of neuronal ensemble inhibition in primary auditory cortex. Neuron, 69 (4):805 -- 817, 2011. I. Ohzawa, G. Sclar, and R.D. Freeman. Contrast gain control in the cat's visual system. Journal of Neurophysiology, 54(3):651 -- 667, 1985. Hyojin Park, Robin AA Ince, Philippe G Schyns, Gregor Thut, and Joachim Gross. Frontal top-down signals increase coupling of auditory low-frequency oscillations to continuous speech in human listeners. Current Biology, 2015. Timothy M Patten, Christopher J Rennie, Peter A Robinson, and Pulin Gong. Human cortical traveling waves: dynamical properties and correlations with responses. PLoS One, 7(6):e38392, 2012. J.E. Peelle and M.H. Davis. Neural oscillations carry speech rhythm through to com- prehension. Front Psychol, 3(320):1 -- 17, 2012. S. Ray, N.E. Crone, . Niebur, P.J Franaszczuk, and S.S. Hsiao. Neural correlates of high-gamma oscillations (60 -- 200 hz) in macaque local field potentials and their potential implications in electrocorticography. The Journal of Neuroscience, 28(45): 11526 -- 36, 2008. D Regan. Some characteristics of average steady-state and transient responses evoked by modulated light. Electroencephalography and clinical neurophysiology, 20(3):238 -- 248, 1966. D. Rubino, K.A. Robbins, and N.G. Hatsopoulos. Propagating waves mediate infor- mation transfer in the motor cortex. Nature neuroscience, 9(12):1549 -- 1557, 2006. T.K. Sato, I. Nauhaus, and M. Carandini. Traveling waves in visual cortex. Neuron, 75(2):218 -- 229, 2012. P. Sauseng, W. Klimesch, W.R. Gruber, S. Hanslmayr, R. Freunberger, and M. Dop- pelmayr. Are event-related potential components generated by phase resetting of brain oscillations? a critical discussion. Neuroscience, 146(4):1435 -- 1444, 2007. C.E. Schroeder and P. Lakatos. Low-frequency neuronal oscillations as instruments of sensory selection. Trends in Neurosciences, 32(1):9 -- 18, 2009. E. Spaak, F.P. de Lange, and O Jensen. Local entrainment of alpha oscillations by visual stimuli causes cyclic modulation of perception. J Neurosci, 34(10):3536 -- 44, 2014. 16 C. Tallon Baudry, O. Bertrand, C. Delpuech, and J. Pernier. Stimulus specificity of phase-locked and non-phase-locked 40 hz visual responses in human. Journal of Neuroscience, 16:4240 -- 4349, 1996. Adriano BL Tort, Robert Komorowski, Howard Eichenbaum, and Nancy Kopell. Mea- suring phase-amplitude coupling between neuronal oscillations of different frequen- cies. Journal of neurophysiology, 104(2):1195 -- 1210, 2010. Bradley Voytek, Andrew S Kayser, David Badre, David Fegen, Edward F Chang, Nathan E Crone, Josef Parvizi, Robert T Knight, and Mark D'Esposito. Oscillatory dynamics coordinating human frontal networks in support of goal maintenance. Nature neuroscience, 2015. E.M. Zion Golumbic, D. Poeppel, and C.E. Schroeder. Temporal context in speech processing and attentional stream selection: a behavioral and neural perspective. Brain and language, 122(3):151 -- 161, 2012. E.M. Zion Golumbic, N. Ding, S. Bickel, P. Lakatos, C.A. Schevon, G.M. McK- hann, R.R. Goodman, R. Emerson, A.D. Mehta, J.Z. Simon, D. Poeppel, and C.E. Schroeder. Mechanisms underlying selective neuronal tracking of attended speech at a "cocktail party". Neuron, 77(5):980 -- 991, 2013. E.M. Zion Golumbic, N. Ding, S. Bickel, P. Lakatos, C.A. Schevon, G.M. McK- hann, R.R. Goodman, R. Emerson, A.D. Mehta, J.Z. Simon, D. Poeppel, and C.E. Schroeder. Mechanisms underlying selective neuronal tracking of attended speech at a cocktail party. Neuron, 77(5):980 -- 991, 2013. 1 A Supplementary Information A.1 Methods A.1.1 Circular statistics concepts This section introduces concepts from circular statistics (Mardia, 1972) used below to define ITC. Given a set of circular variables (e.g., phases), θ1, . . . , θN , we associate to each circular variable a two-dimensional unit vector. Using notation from complex numbers, the unit vector associated with variable θi is: The resultant vector, R, is the sum of the associated unit vectors: vec(θi) = ejθi R(θ1, . . . , θN ) = N X i=1 vec(θi) (1) (2) The mean resultant length, ¯R, is the length of the resultant vector divided by the number of variables: ¯R(θ1, . . . , θN ) = 1 N R(θ1, . . . , θN ) The circular variance, CV , is one minus the mean resultant length: CV (θ1, . . . , θN ) = 1 − ¯R(θ1, . . . , θN ) The mean direction, ¯θ, is the angle of the resultant vector: ¯θ(θ1, . . . , θN ) = arg(R(θ1, . . . , θN )) (3) (4) (5) Note that the mean direction is not defined when the resultant vector is zero, since the angle of the zero vector is undefined. A.1.2 ITC The Inter-Trial Coherence (ITC) is a measure of phase alignment among multiple trials at a given time and frequency (Tallon Baudry et al., 1996; Delorme and Makeig, 2004). Mathematically, it is defined as the mean resultant length ( ¯R, Eq. 3) of the phases of the trials at the given time and frequency. A.1.3 ERSP The Event-Related Spectral Perturbation (ERSP) is the mean power of a set of baseline-normalized epochs (Delorme and Makeig, 2004). 2 A.1.4 Phase-amplitude coupling curve and modulation index The PAC curve is a graphical means of representing the coupling between phases at a modulating frequency and amplitudes at a modulated frequency. The MI is a numerical quantification of the strength of this coupling. We used the methods described in Tort et al. (2010) to compute both the PAC curve and the MI. A.2 Behavioral Data Subjects produced consonant-vowel syllables at different speeds in different experi- mental sessions. This manuscript characterizes data in the experimental sessions with largest and shortest median intersyllable separation times (Figure A.1). A.3 ITC across the vSMC Largest coherence values at the entrained frequency were observed over the vSMC. Figures A.2-A.17 plot the ITC, computed from recordings in experimental session EC2 B105, between the ventral electrode 129 and the dorsal electrode 144. For elec- trodes over the temporal cortex there is almost no ITC (Figures A.2-A.4). ITC in- creases over the ventral section of the vSMC (Figures A.5-A.8), peaks at its center (Figures A.9-A.12), and tapers over its dorsal section (Figures A.13-A.17). 3 0 1 2 3 4 5 6 Inter−Syllable Separation (sec) y c n e u q e r F y c n e u q e r F 0 8 0 6 0 4 0 2 0 0 5 1 0 0 1 0 5 0 0 1 2 3 4 5 6 Inter−Syllable Separation (sec) Figure A.1: Histogram of inter-syllable separations from the experimental ses- sions with the largest (1.62 sec) and shortest (0.97 sec) intersyllable separation time, EC2 B105 (a) and EC2 B89 (b), respectively. 4 Figure A.2: ITC for electrode 129 computed from recordings in experimental session EC2 B105. 5 Figure A.3: ITC for electrode 130 computed from recordings in experimental session EC2 B105. 6 Figure A.4: ITC for electrode 131 computed from recordings in experimental session EC2 B105. 7 Figure A.5: ITC for electrode 132 computed from recordings in experimental session EC2 B105. 8 Figure A.6: ITC for electrode 133 computed from recordings in experimental session EC2 B105. 9 Figure A.7: ITC for electrode 134 computed from recordings in experimental session EC2 B105. 10 Figure A.8: ITC for electrode 135 computed from recordings in experimental session EC2 B105. 11 Figure A.9: ITC for electrode 136 computed from recordings in experimental session EC2 B105. 12 Figure A.10: ITC for electrode 137 computed from recordings in experimental session EC2 B105. 13 Figure A.11: ITC for electrode 138 computed from recordings in experimental session EC2 B105. 14 Figure A.12: ITC for electrode 139 computed from recordings in experimental session EC2 B105. 15 Figure A.13: ITC for electrode 140 computed from recordings in experimental session EC2 B105. A.4 ERSPs across the vSMC Figures A.18-A.30 plot ERSPs at electrodes along the ventro-dorsal axis of the vSMC. Note the sharp transition between the peak of large evoked beta power before the CV transition at electrodes 132-135 over the auditory cortex to the peak of large evoked high-gamma power after the CV transition at electrodes 135-140 over the vSMC. 16 Figure A.14: ITC for electrode 141 computed from recordings in experimental session EC2 B105. 17 Figure A.15: ITC for electrode 142 computed from recordings in experimental session EC2 B105. 18 Figure A.16: ITC for electrode 143 computed from recordings in experimental session EC2 B105. 19 Figure A.17: ITC for electrode 144 computed from recordings in experimental session EC2 B105. 20 Figure A.18: ERSP for electrode 129 computed from recordings in experimental session EC2 B105. 21 Figure A.19: ERSP for electrode 130 computed from recordings in experimental session EC2 B105. 22 Figure A.20: ERSP for electrode 131 computed from recordings in experimental session EC2 B105. 23 Figure A.21: ERSP for electrode 132 computed from recordings in experimental session EC2 B105. 24 Figure A.22: ERSP for electrode 133 computed from recordings in experimental session EC2 B105. 25 Figure A.23: ERSP for electrode 134 computed from recordings in experimental session EC2 B105. 26 Figure A.24: ERSP for electrode 135 computed from recordings in experimental session EC2 B105. 27 Figure A.25: ERSP for electrode 136 computed from recordings in experimental session EC2 B105. 28 Figure A.26: ERSP for electrode 137 computed from recordings in experimental session EC2 B105. 29 Figure A.27: ERSP for electrode 138 computed from recordings in experimental session EC2 B105. 30 Figure A.28: ERSP for electrode 139 computed from recordings in experimental session EC2 B105. 31 Figure A.29: ERSP for electrode 140 computed from recordings in experimental session EC2 B105. 32 Figure A.30: ERSP for electrode 141 computed from recordings in experimental session EC2 B105. 33 Figure A.31: ERP for electrode 129 computed from recordings in experimental session EC2 B105. A.5 ERPs across the vSMC Figures A.31-A.43 plots ERPs computed at electrodes along the ventro-dorsal axis of the recordings grid. Note that ERPs are larger over electrodes 136-140 in the vSMC, in agreement with Figure 2, and that the ERP peak is shifted to later times as we move from electrode 136 to electrode 140. A.6 Electrodes with strongest PAC Figure A.44 highlights the 50 electrodes with largest MI. 34 Figure A.32: ERP for electrode 130 computed from recordings in experimental session EC2 B105. 35 Figure A.33: ERP for electrode 131 computed from recordings in experimental session EC2 B105. 36 Figure A.34: ERP for electrode 132 computed from recordings in experimental session EC2 B105. 37 Figure A.35: ERP for electrode 133 computed from recordings in experimental session EC2 B105. 38 Figure A.36: ERP for electrode 134 computed from recordings in experimental session EC2 B105. 39 Figure A.37: ERP for electrode 135 computed from recordings in experimental session EC2 B105. 40 Figure A.38: ERP for electrode 136 computed from recordings in experimental session EC2 B105. 41 Figure A.39: ERP for electrode 137 computed from recordings in experimental session EC2 B105. 42 Figure A.40: ERP for electrode 138 computed from recordings in experimental session EC2 B105. 43 Figure A.41: ERP for electrode 139 computed from recordings in experimental session EC2 B105. 44 Figure A.42: ERP for electrode 140 computed from recordings in experimental session EC2 B105. 45 Figure A.43: ERP for electrode 141 computed from recordings in experimental session EC2 B105. 46 Figure A.44: Strongest PAC occurs over the vSMC. The 50 electrodes with largest MI (between 0.5 seconds before and 0.6 seconds after the CV transition, at the frequency of CV production, in the experimental session EC2 B105) are highlighted in color. 47
1612.06975
1
1612
2016-12-21T05:30:11
Similarities and differences between stimulus tuning in the inferotemporal visual cortex and convolutional networks
[ "q-bio.NC" ]
Deep convolutional neural networks (CNNs) trained for object classification have a number of striking similarities with the primate ventral visual stream. In particular, activity in early, intermediate, and late layers is closely related to activity in V1, V4, and the inferotemporal cortex (IT). This study further compares activity in late layers of object-classification CNNs to activity patterns reported in the IT electrophysiology literature. There are a number of close similarities, including the distributions of population response sparseness across stimuli, and the distribution of size tuning bandwidth. Statisics of scale invariance, responses to clutter and occlusion, and orientation tuning are less similar. Statistics of object selectivity are quite different. These results agree with recent studies that highlight strong parallels between object-categorization CNNs and the ventral stream, and also highlight differences that could perhaps be reduced in future CNNs.
q-bio.NC
q-bio
Similarities and differences between stimulus tuning in the inferotemporal visual cortex and convolutional networks. Bryan Tripp, University of Waterloo, Canada Submitted on December 1, 2016 to IJCNN 2017. 2016 IEEE. Personal use of this material is permitted. Permission from IEEE must be obtained for all other uses, in any current or future media, including reprinting/republishing this material for advertising or promotional purposes, creating new collective works, for resale or redistribution to servers or lists, or reuse of any copyrighted component of this work in other works. 6 1 0 2 c e D 1 2 ] . C N o i b - q [ 1 v 5 7 9 6 0 . 2 1 6 1 : v i X r a Similarities and differences between stimulus tuning in the inferotemporal visual cortex and convolutional networks. Department of Systems Design Engineering & Centre for Theoretical Neuroscience Bryan P. Tripp Waterloo, Ontario, Canada Email: [email protected] Abstract -- Deep convolutional neural networks (CNNs) trained for object classification have a number of striking similarities with the primate ventral visual stream. In particular, activity in early, intermediate, and late layers is closely related to activity in V1, V4, and the inferotemporal cortex (IT). This study further compares activity in late layers of object-classification CNNs to activity patterns reported in the IT electrophysiology literature. There are a number of close similarities, including the distributions of population response sparseness across stimuli, and the distribution of size tuning bandwidth. Statisics of scale invariance, responses to clutter and occlusion, and orientation tuning are less similar. Statistics of object selectivity are quite different. These results agree with recent studies that highlight strong parallels between object-categorization CNNs and the ventral stream, and also highlight differences that could perhaps be reduced in future CNNs. I. INTRODUCTION Deep convolutional neural networks (CNNs) that are trained for object categorization have activity that resembles that of the ventral visual stream in several respects. For example, early layers often exhibit Gabor-like receptive fields similar to V1 [1], and separation of early layers across multiple GPUs has resulted in separation of color and edge encoding [2], analogous to V1 blobs and interstripes. Furthermore, internal activity in later layers of such CNNs is closely related to activity later in the primate ventral visual stream. [3] showed that much of the variance in activity of IT neurons (up to about 30%) could be accounted for by multilinear regression with the responses of units in the CNN's second-last layers. They also showed that these linear predic- tions had IT-like correlations related to object categories [3]. Models that have been specifically developed to approximate the ventral stream, including HMAX [4] and VisNet [5], have lacked such realistic correlations [6]. [3] also showed that activity in intermediate layers of these CNNs predicted activity of neurons in V4, which is a major source of input to IT [7]. to scale and translation changes [8], and a few are also invariant to ori- entation changes. Consistent with this, [1] found that the vector of population responses in the second-last later of a CNN changed relatively little with with position, size, and orientation, compared to responses in the first layer. Many IT neurons are relatively invariant More recently, [9] found that IT neurons and units of object-classification CNNs provided comparable information about category-orthogonal object properties, including object position and orientation. Importantly, CNNs perform similarly to humans [2], [1] in "core object recognition", a rapid feedforward process that allows primates to recognize objects in natural scenes after seeing them very briefly [10], on similar timescales to inter- saccade intervals. Core object recognition is a sophisticated function, and a major function of the ventral visual stream. It is highly relevant to neuroscience that this this function can now be performed by artificial systems. CNNs can not be considered realistic models of the ventral stream, because they learn in very different ways, consist of much simpler units, typically lack lateral connections, etc. [11]. However, that they have realistic core object recognition capabilities, and also exhibit internal representations that resemble those in the ventral stream, constrains ideas about the roles of these missing details in vision. the fact A. Purpose of this study Linear combinations of CNN activities are currently the best predictors of IT activity in high-variation stimulus condi- tions, but they leave about half the variance of IT responses unaccounted for. It would be interesting to know whether CNNs predict certain factors of variation in IT better than others. Also, the CNN activities themselves (as opposed to optimal linear combinations of them) have somewhat different correlation patterns [3], [6]. The purpose of the present study is to compare the response properties of deep object-recognition CNNs and IT neurons, along several dimensions of stimulus variation that have been explored in IT. Remarkably, despite the optimization of both IT and CNNs for natural scenes, both systems discriminate the same simplified stimuli (e.g. line drawings) in a similar manner (see Results). This study focuses on tuning curves of individual units, allowing more direct comparison with IT than previous investigations of population response vectors [1]. To summarize the results, many response properties of the CNNs are strikingly similar to previously reported responses of IT neurons. These including the distributions of population sparseness across stimuli, and the distribution of size-tuning bandwidth. Other properties, including responses to partial occlusion and clutter, orientation tuning curves, and patterns of size invariance, were somewhat different. Distributions of object selectivities over neurons were quite different. Along with differences in correlations of IT responses within and between object categories [3], [6], these results help to clarify the relationships between stimulus representation in object- classification CNNs and IT. Preliminary versions of some of these analyses appear in [12]. II. METHODS Some of the figures replot responses of IT neurons from the experimental literature. The data points for these plots were extracted from published figures using Web Plot Digitizer (http://arohatgi.info/WebPlotDigitizer/). A. Networks We study responses of two well-known deep convolutional networks. Both networks were trained to classify objects in natural scenes for the ImageNet Large-Scale Visual Recog- nition Challenge [13]. One was Alexnet [2]. This network has previously been compared with IT in [14] and [6]. The other network, VGG-16 [15], has more layers and smaller kernels. Both of the implementations used here were taken from https://github.com/heuritech/convnets-keras. The Alexnet implementation was adapted from a version trained for [16]. The VGG-16 implementation was adapted from a version from https://gist.github.com/baraldilorenzo, which was in turn adapted from the version provided by [15]. Previous comparisons of CNNs with IT [3], [6], [9] have focused on the second-last network layer, a fully-connected layer that precedes the final softmax classification layer. This layer is in a similar position to the inferotemporal cortex (IT) in the ventral stream, the neurons of which are important for visual object categorization, but also contain much non- category information, e.g. [8], [9]. For comprehensiveness, we also consider responses from the surrounding layers, i.e. the last and third-last layers, which turn out to more closely resemble IT in some respects. B. Stimulus Images Stimulus images were obtained from several sources. To investigate stimulus selectivity and population response sparse- ness, we used a set of 806 images developed by [17] and used for the same purpose in IT. The images were ex- tracted directly from their supplementary materials file. The extracted images had a lower resolution (60 by 60 pix- els) than the authors reported using (125 by 125 pixels). However, the objects were easily identifiable. The extracted images were embedded within a larger gray background. The extracted images themselves had light borders and some- what noisy gray backgrounds. The borders were removed by omitting the outer two pixels, and variations were re- duced in the backgrounds by setting to a uniform gray value ([red,green,blue]=[128, 128, 128]T) any vector with a cartesian distance of less than 15 from this value. A few additional higher-resolution images (banana, car, shoe, etc.) were downloaded from various internet sources and given uniform gray backgrounds. These images were used to examine orientation, position, and size tuning. Images for studying responses to rotation-in-depth were rendered in SketchUp. 3D scooter and human head models were obtained from the SketchUp 3D Warehouse. A wire-like object model, as in [18] was built in OpenSCAD. Other images were generated using custom code in order to approximate stimuli in [19]. Finally, some stimulus images were copied from figures in [8] and [20]. III. RESULTS A. Stimulus selectivity and population sparseness Two fundamental statistical properties of IT population responses are 1) the sparseness of responses across the pop- ulation to various stimuli, and 2) the sparseness of individual neurons' responses evaluated over multiple stimuli (sometimes called stimulus selectivity). Both have been described in terms of excess kurtosis (kurtosis in excess of that of the Gaussian distribution). [17] compared stimulus selectivity and population sparseness in a large dataset that included responses of 674 IT neurons to 806 different images. They found that distributions of selectivity and sparseness had a similar shape, and that sparseness exceeded selectivity. They found that this difference could be reproduced in a simple statistical model, as a result of non-uniform mean spike rates across the population. Datasets similar to that of [17] were collected from the CNNs, using approximately the same set of images (see Methods). Figure 1 plots selectivity vs. sparseness of units in the final three layers of both CNNs (Alexnet and VGG-16). 674 units were chosen from each layer, to match the size of the dataset in [17]. The 674 units with the highest peak activations in response to the stimulus images were chosen from each layer. The sparseness of the CNN activity was similar to that of the IT neurons. However, in contrast to IT neurons, each of the CNN layers exhibited higher selectivity than sparseness. The sparseness of the CNN activity was disproportionately due to very rare activation of some units, while the remaining units were activated less sparsely than average. The sparseness distributions had a similar shape to that of [17], but the selectivity distributions had longer tails (not shown). B. Orientation Tuning The responses of most IT neurons varies strongly with orientation in each dimension [18], although a small fraction (< 1◦ in [18]) are essentially orientation-invariant. Responses are typically strongest around a single preferred orientation unless the object is rotationally symmetric. Orientation tun- ing curves often resemble gaussian functions. [18] reported somewhat narrower tuning curves for rotation in depth than rotation in the image plane. For neurons that responded to faces, [21] reported a fairly uniform distribution of preferred viewing angles between left and right profiles (their Figure 11). s s e n e s r a p S 35 30 25 20 15 10 5 0 0 IT out-1 out out-2 out out-1 out-2 5 10 15 20 25 30 35 Selectivity Fig. 1. Sparseness vs. selectivity of IT neurons (from [17]), and of CNN units responding to a similar group of stimuli. Sparseness exceeds selectivity in IT [17], whereas selectivity consistently exceeds sparseness, in each layer of both CNNs. The blue circles are Alexnet layers and the green squares are VGG-16 layers. The labels indicate the distance from the output, i.e. "out" is the last layer, "out-1" is the second-last, and "out-2" is the third-last. Figure 2 shows tuning curves over rotations in depth about the vertical axis. Figure 2A shows example views of three stimulus objects: a "wire-like" object, similar to those used in [18]; a more complex inanimate object that might be encountered in life (a scooter); and a human head. Figure 2B shows examples of IT-neuron orientation-tuning curves for a wire-like object (replotted from [18]) and a human head (replotted from [22]). The wire (left) elicited narrow, roughly Gaussian tuning curves from most neurons [18]. A minority of neurons responded strongly to roughly mirror-symmetric views separated by 180deg. Finally, a very small minority of neurons had approximately view-invariant responses. The left columns of Figure 2C and D show responses of units in Alexnet and VGG-16, respectively. In contrast with IT neurons, the CNN units typically have multimodal tuning curves. Units in the last two layers have broad tuning. However, the third-last layer of Alexnet has several fairly narrow, roughly Gaussian tuning curves that resemble view- selective neurons in [18], except that they are bimodal (perhaps due to use of flipped images to enhance the dataset in training). As the wire-like object is unlike objects that the networks encountered in training, orientation tuning curves were also plotted for an object from one of the ImageNet categories (a scooter). As with the wire, most responses in both Alexnet and VGG-16 were broad and multimodal. There are no examples of narrow uni- or bimodal Gaussian tuning in this column. The right column of 2B shows responses of several cells in face-selective regions of IT from [22]. Some of these cells preferred front views, and some preferred profile views, but responded similarly to left and right profiles. Interestingly, CNN tuning curves also showed these two patterns, although with somewhat broader tuning. In [18], rotations in the image plane produced responses that were similar to rotation in depth (many neurons had roughly Gaussian tuning) except that tuning was somewhat broader. 8 7 6 5 4 3 2 1 0 1 0 50 100150200250300350 6 5 4 3 2 1 0 1 0 50 100150200250300350 e s n o p s e R e s n o p s e R   5 4 3 2 1 0 1 0 50 100150200250300350 6 5 4 3 2 1 0 1 0 50 100150200250300350  Angle (degrees) Angle (degrees) Fig. 3. Tuning curves for image-plane orientation of staple, shoe, car, and banana images. In each case, the responses of the first ten units with peak responses > 2 are shown. The plotted tuning curves were smoothed by averaging over groups of four neighbouring points, so that the different curves are easier to distinguish, according to the Gestalt law of good continuation. Figure 3 shows examples of image-plane orientation tuning curves from the second-last layer of Alexnet. In contrast with typical IT responses to image-plane rotation, most of the CNN responses are multimodal. C. Size Tuning and Invariance IT neurons are also tuned for size. [23] studied the distri- bution of size tuning bandwidths, i.e. the ratio between the larger and smaller sizes at which the response drops to half it's peak. The mode of this distribution was approximately 1.5 octaves (their Figure 8), and bandwidths of up to 4 octaves were reported. Figure 4B and C shows histograms of size tuning bandwidths in Alexnet and VGG-16, respectively. They are both quite similar to the distribution in IT. For example the Alexnet histogram has a mode around 1 (versus 1.5 in IT), and values up to 4.6 (vs. a few values at or above 4, the highest value measured, in IT). Responses of the majority of IT neurons exhibit size in- variance [8]. Spike rates vary with stimulus size, but the relative spike rates in response to different stimuli are similar regardless of sizes. A prototypical example is shown in Figure 5A. The neuron's tuning curve over six different shapes is roughly maintained over different stimulus scales. However, the gain of this curve depends on the scale. Figure 5D shows responses of units in the second-last layer of Alexnet. The stimulus shapes and relative scales are the same as in [8]. Similar to the IT neurons, many of the CNN units' tuning curves have similar shapes when the stimuli are shown at different scales. A 'c' in the upper-right of each plot indicates A B C D s n o r u e N T I 60 50 40 30 20 10 e s n o p s e R 0 180  e s n o p s e R e s n o p s e R e s n o p s e R e s n o p s e R e s n o p s e R e s n o p s e R t e n x e A l 6 1 - G G V 180 4 3 2 1 0 1 2   5 4 3 2 1 0 1 180 " 25 20 15 10 5 0 5 180 * ( 180 4 3 2 1 0 1 24 3 1 7 6 5 4 3 2 1 0 1 180 A ? 16 14 12 10 8 6 4 2 0 2 180 J H Staple Scooter Head 0° 30° 0° 30° 0° 30° E 18 16 14 12 10 8 6 4 2 0 180  180 60 60  Angle (degrees) 180 60 60  Angle (degrees) 3.0 2.5 2.0 1.5 1.0 0.5 0.0 0.5 1.0 1.5 180  8 7 6 5 4 3 2 1 0 1 180 % # 25 20 15 10 5 0 5 180 - + 180 5 4 3 2 1 0 1 29 38 7 5 10 8 6 4 2 0 2 180 D B 20 15 10 5 0 5 180 M K 60  60 180    60 ! 60 180 60 60 180 ) Angle (degrees) 60 2 60 180 60 @ 60 180 60 60 180 I Angle (degrees) 4.0 3.5 3.0 2.5 2.0 1.5 1.0 0.5 0.0 0.5  180 8 6 4 2 0 180 & 30 25 20 15 10 5 0 5 180 0 . 180 5 4 3 2 1 0 1 2> 3= < : 5 4 3 2 1 0 1 180 G E 12 10 8 6 4 2 0 2 180 P N 60  60 180  60 $ 60 180 60 60 180 , Angle (degrees) 60 6 60 180 60 C 60 180 60 60 180 L Angle (degrees) 60  60 180 F 60 ' 60 180 60 60 180 / Angle (degrees) 60 ; 60 180 60 F 60 180 60 60 180 O Angle (degrees) s l l e c f o n o i t c a r F s l l e c f o n o i t c a r F s l l e c f o n o i t c a r F s t i n u f o n o i t c a r F s t i n u f o n o i t c a r F s t i n u f o n o i t c a r F AM 0.5 0.4 0.3 0.2 0.1 0.5 0.0 0.5 1.0 AL 0.5 0.0 0.5 1.0 ML/MF 1.0 0.0  0.5 0.4 0.3 0.2 0.1 1.0 0.0 0.5 0.4 0.3 0.2 0.1 0.0 1.0 0.5 0.0 0.5 1.0 Head orient at ion t uning dept h output 0.5 0.4 0.3 0.2 0.1   0.5 0.0 0.5 1.0 output-1 0.5 0.0 0.5 1.0 output-2 1.0 0.0  0.5 0.4 0.3 0.2 0.1 1.0 0.0  0.5 0.4 0.3 0.2 0.1 0.0 1.0 0.5 0.0 0.5 1.0  Head orient at ion t uning dept h  Fig. 2. Tuning curves for rotation-in-depth of a wire-like object (first column), a scooter (second column), and a human head (third column). A) Example views that were used as inputs to the CNNs. B) Tuning curves of IT neurons in response to similar (not identical) objects. Responses to a wire-left object (left) and head (right) are redrawn from data in [18] and [22], respectively. C) Responses of units in the final three layers of Alexnet. The columns correspond to wire, scooter, and head stimuli. The top row shows responses from the output layer, and the other rows show responses from the immediately preceeding layers. In each case the responses of the first ten units with peak responses greater than two are shown, and the tuning curves are smoothed by averaging over groups of three neighbouring points. D) Same format as C for VGG-16. E) Histograms of tuning depth of IT neurons in different face-responsive areas, replotted from [22]. The tuning depth is the frontal response minus the profile response, divided by their sum. Depths are closer to zero (tuning curves are flatter) in the anterior medial area (AM) relative to more posterior areas. F) Depth histograms of head tuning curves from the final three layers of Alexnet. Analogous to the neural data, depth is lowest in the output layer. In contrast with the neural data, the distributions in other layers are multimodal. A y c n e u q e r F D 9 8 7 6 5 4 3 2 1 0 0 1 2 3 4 5 6 Size BW (Oct aves) B 140 120 100 80 60 40 20 0 0 1 2 3 4 5 6 y c n e u q e r F C 160 140 120 100 80 60 40 20 0 y c n e u q e r F 0 1 2 3 4 5 6 Size BW (Oct aves) Size BW (Oct aves) 250 200 150 100 50 0 − 50 − 100 1 A ) s / s e k i p s ( e s n o p s e R D largest B 1.0 0.8 0.6 0.4 0.2 n o i t a l e r r o Q n a e M Alexnet VGG-16 smallest 4 3 2 St im ulus # 5 6 0.0 − 2 Dist an c − 1 0 e from out put (layers) C e nT e r e f e r p h t i w n o i t aS r F 1.0 0.8 0.6 0.4 0.2 0.0 − 2 Dist an R − 1 0 e from out put (layers) Fig. 4. A) Histogram of half-magnitude size tuning bandwidths, replotted from [23]. The range of stimulus scales did not fully span the half-magnitude ranges of the units plotted in red. B) Histogram of half-magnitude bandwidths for Alexnet. Stimuli included scaled versions of banana, shoe, and car images. For each object, the 200 units with the strongest peak responses were included in the histogram. C) As B, but for VGG-16. D) Examples of banana images at some of the different scales used with the CNNs. that the mean correlation between curves is greater than 0.75. However, only a small minority of CNN units have a clear preferred size (marked with 'p' in the upper left) like the canonical example of [8]. Panels 5B and C show that these results are similar for other layers of Alexnet and VGG-16. Mean correlations are well below those of 5A, and only a small fraction of CNN units in any layer show a clear preference for a certain size that is maintained across multiple shapes. Shape and size tuning may be somewhat less separable in the CNNs than in IT. D. Position Tuning and Invariance Many IT neurons exhibit translation invariance as well as size invariance [23]. Figure 6A shows examples of response variations with horizontal stimulus position, in the second- last layer of Alexnet. In each plot, the first 30 units with responses > 2 are shown. As in [24], many of the tuning curves are roughly gaussian, and most units prefer centrally located stimuli. The dispersion of preferences around zero (the fovea) cannot be directly compared, because the mapping between pixels and degrees visual angle is not well defined (it is not uniform in the ImageNet training data). However, it is possible to compare the ratio of this dispersion to the widths of the tuning curves. Figure 6B plots ratios of standard deviation of tuning curve centres of mass, divided by mean tuning curve width (distance between points on each side of the centre of mass that fall below half the peak). The dashed line is this ratio taken from the IT data in [24]. The blue and green lines are from Alexnet and VGG-16, respectively. As a fraction of tuning curve width, the position-tuning centres in the last layers of both networks are much more narrowly dispersed than IT position-tuning curves. However, the second-last layer of Alexnet and third- last layer of VGG-16 are similar to the IT neurons in this respect. Figure 7 shows that (in contrast with size invariance) the CNNs have similar position invariance to IT. The last and c c c c c c c p c c p c c c c c c c p c c c c c c c c Fig. 5. Size invariance in the CNNs, examined with stimuli from [8]. A) Responses of an example IT neuron, replotted from [8]. The horizontal axis spans different stimulus shapes, the vertical axis is the mean spike rate. The red, green, and blue curves correspond to small, medium, and large versions of the stimuli. B) The dashed line is the correlation between the tuning curves in A. The blue and green are mean correlations of similar tuning curves in Alexnet and VGG-16. The correlations are plotted versus layer, with the output layer on the right. C) The fraction of units in each CNN layer that had the same scale preference for all shapes (like the example neuron in A, in which the red curve does not cross the others). D) Examples of tuning curves in the second-last layer of Alexnet, with the same format as A. Many of the units' responses are correlated across scales (those with mean correlations ¿0.75 are marked 'c'). However, few have a uniform scale preference across shapes (marked 'p'). The units in B-D were the first 64 units in each layer with peak responses > 2 over all stimuli. second-last CNN layers have average correlations in these tuning curves that are similar to the example given by [8] (their Figure 2D), which is consistent with the population averages of [24] (their Figure 10C). A substantial fraction of the units have a position preference that is consistent across shapes, with at most one shape at which the preferred position is inconsistent (as in the IT data in 7A). E. Occlusion [19] studied variations in the responses of IT neurons as shape stimuli were partially occluded. Neurons maintained their shape preferences, but responded less vigorously as the shapes were occluded more fully. Figure 8 shows results of a similar experiment with the CNN. For this experiment, following [19], several shape outlines were drawn on top of a noisy background, and these shapes were occluded by replacing blocks of pixels with black. As with IT neurons, shape preference order tended to be maintained with increasing occlusion, and the amplitude of the preferences was reduced with high occlusion. 8A shows A 6 5 4 W 2 1 0 − 1 − 8 6_ 5 4 2^ 1 0 − 1 − [ { st aple shoe 5 4 Z 2 1 0 5 − 25 25 U 5 V − 1 − X 5 − 25 25 5 Y ar ` 5 − 25 25 ffset (pi els) x e s n o p s e R 6g 5 4 2f 1 0 − 1 − 5 \ anana h a 5 − 25 25 ffset (pi els) e d 5 b ] h t io w n a e m / s r e t n e f n o Alexnet 0.5 0.4 0. j 0.2 0.1 IT ons pqrs tuvwyz 0.0 − 2 − 1 0 Dist an i e from out put (layers) 400 50 00„ 250ƒ 200 150 100 50 0 − 50 1 A ) s / s e k i p s ( e s n o p s e R « p 2 ˆ Ž ˜ p p © 1.0 n o i t a l e r r o € n a e M 0.8 0.6 0.4 0.2 ¬-®¯°± ²³´µ¶· ¸¹º»¼ ½¾¿ÀÁ ÂÃÄÅ VGG-16 Alexnet 4 5 6  t im ulus # ‚ … p ’ ž ¡ 0.0 − 2 Dist an  − 1 0 e from out put (layers) Š  “ š p p ‹ ” › Ÿ p p ‰  ™ ¢ ¥ e ª n~ e r e f e r p h t i w n o i t a} r F 1.0 0.8 0.6 0.4 0.2 Alexnet VGG-16 − 1 0 e from out put (layers) 0.0 − 2 Dist an † • ¦ Œ – œ £ § ‡  ‘ —  ¤ Fig. 6. A) Responses to staple, shoe, car, and banana images translated to different horizontal positions within a gray background. The offset is the distance from centre, in an image that was 256 pixels wide. At the largest offsets, the stimuli nearly reached the edge of the background. B) Ratios of standard deviation of centres of mass, over mean half-width, of position- tuning curves. The dashed line is the ratio from data in [24]. The blue and green lines are calculated from the responses of Alexnet and VGG-16 to the translated staple, shoe, car, and banana images. Fig. 7. Translation invariance in CNN units. Both the plot format and CNN unit selection procedure are the same as in Figure 5. A) Data replotted from an example neuron in [8]. B) Mean correlations of tuning curves in Alexnet (blue) and VGG-16 (green) compared with that of the curves in panel A (dashed). C) Fraction of units in each layer in which the same position was preferred for all but at most one shape, like the IT neuron shown in panel A. D) Example tuning curves from the second-last layer of Alexnet, in the same format as A. The different tuning curves correspond to stimuli at the centre, and translated up, down, left, and right from centre. The translation distance relative to stimulus size was roughly matched to that in [8]. tuning curves over different shapes, with different occlusion levels, for the second-last layer of Alexnet (left), the second- last layer of VGG-16 (centre), and IT neurons from [19]. The stimuli used to generate these plots were black outlines, and the occluders were black squares (see example in Figure 8B). In these examples, partial occlusion attenuated the CNN unit responses more strongly than the IT responses. For example, at 50% occlusion, VGG-16 units responded uniformly and minimally regardless of the stimulus shape. Figure 8B plots mean normalized responses to units' pre- ferred shapes, as a function of the unoccluded percentage of the image. Compared to the IT neurons (black line), responses in each layer of Alexnet (left) and VGG-16 (right) are attenuated much more by occlusion. The CNN and IT responses cannot be compared directly, due to temporal evolution of the IT stimulus over time that cannot be reproduced in the CNNs. For example, in the IT experiment, the shape appeared after the occluder, providing an additional segmentation cue. To provide a comparable cue for the CNNs, the simulations were repeated with the shapes and occluders in different colors (right panel of 8C). This reduced the attenuation of CNN responses with occlusion. Another complication is that the IT data shown in 8A (right) is from experiments in which the occlusion patterns moved. To approximate such motion, the CNN experiment was repeated with occluders that were blurred over a similar distance relative to the stimulus sizes (right panel of 8D). This stimulus change also reduced the attenuation of CNN responses with occlusion. In this condition, Alexnet's occlusion responses were similar to those of IT neurons. Notably however, the IT neurons responded similarly to 50% occlusion regardless of whether it was moving or stationary [19]. Relatedly, occlusion of certain parts of objects impairs the performance of CNNs more than occlusion of other parts [1], consistent with responses in IT [25]. F. Clutter [26] compared the responses of IT neurons to isolated stimulus objects with responses to multiple objects. Responses to pairs of objects were closely clustered around the average response (rather than the sum) of responses to each object individually 9A. A similar experiment was performed with the CNN, using eights images from the dataset of [17]. Images were shown at two locations (above and below centre) in a gray background, as in the example of Figure 9D. Four images were used in the upper location, and four different images were used in k l m A e s n o p s e R n a e M 1.6 1.4 1.2 1.0 0.8 0.6 0.4 0.2 0.0 − 0.2 1 2 Alexnet VGG-16 IT Neurons A ÐÑ20 Ò50 Ó 0 Õ Ô e s n o p s e R n a e M 4.0 .5 .0 Ù Ø 2.5 2.0 1.5 1.0 0.5 5 4 6 Î Shape rank 0.0 1 2 8 Ï ÚÛ20 Ü50 Ý 0 ß Þ 25 20 15 10 5 0 e s n o p s e R n a e M − 5 1 2 8 × ÈÉ20 Ê50 Ë 0 Í Ì 8 Ç 5 4 6 AE Shape rank 5 4 6 Ö Shape rank B C D e s n o p s e R l eae iå a m r o ä n a e M e s n o p s e R l eò iñ a m r o ð n a e M e s n o p s e R l eþ iý a m r o ü n a e M 1.0 0.8 0.6 0.4 0.2 0.0 1.0 0.8 0.6 0.4 0.2 0.0 1.0 0.8 0.6 0.4 0.2 0.0 çèlast layer last 2n é last ë r ê 20 40 60 er à á ent isi â 80 le ã 100 óôlast layer last 2n õ last ÷ r ö 20 40 60 er ì í ent isi î 80 le ï 100 ÿTlast layer last 2n d last r 3 20 40 60 er ø ù ent isi ú 80 le û 100 e s n o p s e R l e iz a m r o N n a e M e s n o p s e R l e i a m r o n a e M e s n o p s e R l e i a m r o  n a e M 1.0 0.8 0.6 0.4 0.2 0.0 1.0 0.8 0.6 0.4 0.2 0.0 1.0 0.8 0.6 0.4 0.2 0.0 Ilast layer last 2n  last  r  20 40 60 er P c ent isi V 80 le b 100 last layer last 2n  last  r  20 40 60 er   ent isi  80 le 100 last layer last 2n  last  r  20 40 60 er   ent isi  80 le  100 Fig. 8. Results of an occlusion experiment. A) Shape tuning curves averaged over the 500 units with the strongest responses from the second-last layers of Alexnet (left) and VGG-16 (center), and data from IT neurons in [19] (right). The CNN responses are averages over ten random occlusion patterns (i.e. different random placements of occluding blocks). An example stimulus is shown on the right of panel B. The background, shapes, and occluders were intended to approximate those in [19]. The lines correspond to no occulusion, and occlusion of 20%, 50%, and 90% of the image. These responses have baseline activity subtracted. In [19], baseline activity was taken from an inter- stimulus delay period. In the CNNs, baseline activity is the response to a black image. B) Responses to preferred stimuli at different occlusion levels. The black, blue, and green lines are IT data from [19], and unit responses from Alexnet and VGG-16, respectively. These responses were normalized by subtracting the mean response to the least-preferred stimuli, and dividing by the response to the unoccluded preferred stimuli. An example star-shaped stimulus with 20% occlusion is shown on the right. C) As B, but with stimulus shapes in red rather than black. D) As B, but with the occlusion patterns blurred to approximate motion of the occluder in [19]. IT ons #$%& B s r i a p t e" o o t e s n o p s e R 16 14 12 10 8 6 4 2 0 0 Ale x'() *+ d-Last La y,- 2 4 6 8 10 12 14 16 Sum of responses t o sin le o t s e  j g C Alexnet D output output-1 output-2 0 pi/4 0 pi/4 0 pi/4 VGG-16 output output-1 output-2 pi/4 0 Angle of (sum, pair) point from horizontal pi/4 pi/4 0 0 Fig. 9. Effects of low levels of clutter. A) Responses of IT neurons to pairs of objects (vertical axis) versus sums of their responses to each object presented alone (reproduced with permission from [26]). Red dots indicate that each stimulus elicited a response greater than the background rate, and blue dots indicate that only one stimulus did so. The responses to pairs of objects are mostly close to the average of responses to each object alone. B) Responses of units in the second-last layer of Alexnet to pairs of objects versus sums of responses to each object alone. Compared to IT, the CNN responses to paired objects are more varied, and are typically greater than the mean of the single-object responses. The plot includes responses of the 100 units with the greatest peak responses across pair conditions. A random selection of units produced a similar pattern but with a higher density of points in the lower left corner. C) Angles of points (counter-clockwise from rightward) in the same space as the scatterplots in panels A an B. Each histogram corresponds to a layer of one of the CNNs. The red lines indicate the angle arctan(1/2). This corresponds to points on the solid lines in A and B, i.e. responses to paired objects that equal the average of responses to individual objects. The average angle for the IT responses in A is approximately arctan(1/2). D) An example of a pair of images. of π/4 means that the unit's response to paired images equals the sum of responses to each image alone. An angle of arctan(1/2) (red lines) means that the unit's response to paired images equals the average of responses to each image alone. In each layer of both networks, the mean and mode of the angles is between arctan(1/2) and π/4, suggesting that the CNN responses are not normalized as strongly as IT responses. This is true even in the output layers, which have softmax nonlinearities. G. Strong Responses to Simplified Images the lower location. Unit responses were recorded with each of the upper and lower stimuli alone, and also with each of the 16 possible pairs of upper and lower stimuli. As shown in 9B, responses of the second-last layer of Alexnet to pairs of images did not cluster tightly around the average of responses to corresponding isolated images (in contrast with [26]), but were more broadly distributed, and typically intermediate to the sum and mean of the isolated responses. Figure 9B shows histograms of the angles of these points, counter-clockwise from the positive horizontal axis. An angle Tanaka and colleagues found that many IT neurons that responded strongly to complex images also responded strongly to highly simplified versions of these images [20]. After showing a monkey many objects in a standard set, and determining the most effective stimulus from the set (i.e. the one that elicited the strongest response), the most effective stimulus was progressively simplified to determine which features were necessary for maximal activation. For example, a cell that responded maximally to a sketch of a striped cat also responded maximally to a sketch of a striped ball on ! top of another striped ball. Response strength often increased with simplification. IT neurons typically responded maximally to moderately complex stimuli that were more complex than primitive shapes, but less complex than natural objects. The image-simplification process was manual and not fully specified, so it is not clear how accurately it could be re- produced for CNN units. Instead, the CNNs were tested with twelve examples of complex and simplified images provided in [20]. The CNN was first shown the twelve complex images and 56 additional images taken from ShapeNet (total 68 images, similar to the numbers of stimuli described in [27]). For each of the 12 complex images, units were identified that responded maximally or nearly maximally (within 5% of the mean unit- wise maximum) to that image among the 68 stimuli. Figure 10A shows distributions of response strengths of these units when shown the corresponding simplified images, normalized to the response response to the complex image. The responses to simplified images are broadly distributed, but in many cases, a substantial fraction of the units responded at least as strongly to the simplified image as to the complex image. An exception was that very few of the units that responded strongly to a person wearing a lab coat responded as strongly to its simplification (a black circle over a white circle). This exception was consistent across other Alexnet layers, and across the VGG-16 layers. There are infinite ways in which an image could potentially be simplified, so it is striking that the particular simplifications that elicited maximal responses from the IT neurons in [20] also elicited maximal responses from some of the 4098 units in this layer of Alexnet. Figure 10B shows the fractions of neurons in each layer of each network that responded maximally to the simplified images in [20]. Interestingly, a few of the units in each layer responded maximally to simplifications of other objects (e.g. a unit that responded maximally to a complex bread image may also have responded maximally to a simplified cat). However, the fractions of maximal responses to simplification of the most effective stimulus were uniformly greater. IV. DISCUSSION Past studies [1], [3], [6], [9] have shown that activity in the next-to-last layer of object-classification CNNs is closely related to that in the primate inferotemporal cortex (IT). The current study examined properties of the tuning of individual neurons to a number of stimulus variations that have been used to understand representation in IT. The results highlight some specific similarities between the CNN units and IT neurons, including particularly the distributions of population response sparseness, and size-tuning bandwidth. A number of other tuning properties had both similarities and differences with IT. These include scale and translation invariance, orientation tuning, and responses to occlusion. Responses to clutter were also systematically different. The distributions of stimulus selectivity were quite different. Overall, in agreement with past studies, the similarities are striking, particularly in light of the han : 1 2 skew er 1 2 rea < ; 2 1 at = all 7 1 2 hayst a k 8 1 2 pom pom 9 1 2 apple 0.8 0.4 0.0 − 0.4 − 0.8 0 0.8 0.4 0.0 − 0.4 − 0.8 0 0.8 0.4 0.0 − 0.4 − 0.8 0 0.8 0.4 0.0 − 0.4 − 0.8 0.8 0.4 0.0 − 0.4 − 0.8 0 0.8 0.4 0.0 − 0.4 − 0.8 0 0.8 0.4 0.0 − 0.4 − 0.8 0 0.8 0.4 0.0 − 0.4 − 0.8 0 A 0.8 0.4 0.0 − 0.4 − 0.8 0 0.8 0.4 0.0 − 0.4 − 0.8 0 0.8 0.4 0.0 − 0.4 − 0.8 0 s t i n aJ r F plant 1 2 hin le 4 2 1 2 e u 5 6 1 2 person 0.8 0.4 0.0 − 0.4 − 0.8 R 1 0 1 atio of response to simple 2 2 0 1 2 x stimulus >?@ BCDEFH B 1 > o i t aW r F 16 .G/0 0.20 0.15 0.10 0.05 le 1 A net 0.20 0.15 0.10 0.05 0.00 − 2.0 − 1.5 − 1.0 − 0.5 0.0 0.00 − 2.0 − 1.5 − 1.0 − 0.5 0.0 Distan ^_ from output ( l`aefhi Fig. 10. Responses of CNN units to simplified versions of most effective stimuli. A) Responses of the second-last layer of Alexnet. Each histogram corresponds to an example pair of complex and simplified images from Figure 1 of [20]. Neurons were found that responded maximally to each of the complex stimuli. The blue histograms show the distribution of strength of responses to the corresponding simplified stimuli (normalized by strength of response to the most effective complex stimulus). As a control, the gray histograms show the distribution of strength of responses to other (non-matching) simplified stimuli. B) Fractions of neurons with equal or greater responses to simplified stimuli than complex stimuli. The blue lines correspond to the matching simplified stimulus for each complex stimulus, and the gray lines correspond to the other simplified stimuli. many differences in computational mechanisms and details of learning between the visual cortex and the CNNs. Others [3], [14], [6] have found relationships between IT activity and CNNs. However, they did not report individual neuron tuning properties, so similarity with IT may have varied differently with respect to different tuning properties in these networks. A. Limitations This study has examined a small selection of tuning prop- erties, and the responses of only two CNNs, although these CNNs are widely used and reasonably representative. A more comprehensive investigation would examine the effects of ad- ditional CNN structures and training procedures. Furthermore, although efforts were made to approximate the stimuli used in IT studies, the stimuli could not typically be reproduced exactly. Generally an exact analogy isn't possible. For exam- ple, the number of pixels per degree visual angle is not well defined in the networks. Also, the stimuli in the IT studies K L M O Q S U X Y Z [ \ ] appeared and disappeared, and sometimes moved, and the IT neurons responded dynamically. In contrast, the CNNs receive single frames of input and produce a single corresponding response (without dynamics). CNNs responses are probably most closely related to the earliest (largely feedforward) IT responses. However, IT responses change constantly, and latencies vary across conditions, so it is not clear which time window of IT responses would be the most closely related. B. Toward IT-Like CNNs The many similarities between CNNs and the ventral visual stream suggest an opportunity to further optimize CNNs to more closely approximate the ventral stream, perhaps in terms of size invariance and orientation tuning. [3] trained CNNs specifically to emulate IT neuron activity, but this approach did not yield better IT approximations than simply training the CNNs for object classification. However, this attempt may have been hampered by the small size of the neural dataset. Another approach would be to use an empirical model of cortical activity to produce a large (but approximate) set of labels, as in [28]. Further incorporation of physiological mechanisms such as [29] into CNNs may also be important. V. CONCLUSION This work is a preliminary comparison of various stimulus tuning properties of CNN units to corresponding properties of IT neurons. The results show a full spectrum of close simi- larities, subtle differences, and large differences with respect to different tuning properties. This provides some missing detail on the specific relationships between CNN unit and IT neuron responses, and suggests how CNN unit activity could be modified in future work to more closely reflect the visual representation in IT. ACKNOWLEDGMENT Victor Reyes Osorio extracted the stimulus images from [17]. Salman Khan provided comments on an earlier draft. REFERENCES [1] M. D. Zeiler and R. Fergus, "Visualizing and Understanding Convolu- tional Networks arXiv:1311.2901v3 [cs.CV] 28 Nov 2013," Computer VisionECCV 2014, vol. 8689, pp. 818 -- 833, 2014. [2] A. Krizhevsky, I. Sutskever, and G. E. Hinton, "ImageNet Classification with Deep Convolutional Neural Networks," in Advances in Neural Information Processing Systems, 2012, pp. 1 -- 9. [3] D. L. K. Yamins, H. Hong, C. F. Cadieu, E. a. Solomon, D. Seibert, and J. J. Dicarlo, "Performance-optimized hierarchical models predict neural responses in higher visual cortex." PNAS, may 2014. [4] T. Serre, A. Oliva, and T. Poggio, "A feedforward architecture accounts for rapid categorization." PNAS, vol. 104, no. 15, pp. 6424 -- 9, apr 2007. [5] E. T. Rolls, "Invariant Visual Object and Face Recognition: Neural and Computational Bases, and a Model, VisNet." Frontiers in Computational Neuroscience, vol. 6, no. June, p. 35, jan 2012. [6] S. M. Khaligh-Razavi and N. Kriegeskorte, "Deep Supervised, but Not Unsupervised, Models May Explain IT Cortical Representation," PLoS Computational Biology, vol. 10, no. 11, 2014. [7] N. T. Markov, M. M. Ercsey-Ravasz, a. R. Ribeiro Gomes, C. Lamy, L. Magrou, J. Vezoli, P. Misery, A. Falchier, R. Quilodran, M. a. Gariel, J. Sallet, R. Gamanut, C. Huissoud, S. Clavagnier, P. Giroud, D. Sappey- Marinier, P. Barone, C. Dehay, Z. Toroczkai, K. Knoblauch, D. C. Van Essen, and H. Kennedy, "A weighted and directed interareal connectivity matrix for macaque cerebral cortex." Cerebral Cortex, vol. 24, no. 1, pp. 17 -- 36, jan 2014. [8] E. L. Schwartz, R. Desimone, T. D. Albright, and C. G. Gross, "Shape recognition and inferior temporal neurons," PNAS, vol. 80, no. September, pp. 5776 -- 5778, 1983. [9] H. Hong, D. L. K. Yamins, N. J. Majaj, and J. J. DiCarlo, "Explicit information for category-orthogonal object properties increases along the ventral stream," Nature Neuroscience, vol. 19, no. 4, pp. 613 -- 622, 2016. [10] J. J. DiCarlo, D. Zoccolan, and N. C. Rust, "How does the brain solve visual object recognition?" Neuron, vol. 73, no. 3, pp. 415 -- 434, 2012. [11] L. Robinson and E. T. Rolls, "Invariant visual object recognition: biologically plausible approaches," Biological Cybernetics, vol. 109, no. 4-5, pp. 505 -- 535, 2015. [12] S. Khan and B. Tripp, "An empirical model of activity in macaque inferior temporal cortex," Neural Networks, accepted. [13] O. Russakovsky, J. Deng, H. Su, J. Krause, S. Satheesh, S. Ma, Z. Huang, A. Karpathy, A. Khosla, M. Bernstein, A. C. Berg, and L. Fei-Fei, "ImageNet Large Scale Visual Recognition Challenge," International Journal of Computer Vision, vol. 115, no. 3, pp. 211 -- 252, 2015. [14] C. F. Cadieu, H. Hong, D. L. K. Yamins, N. Pinto, D. Ardila, E. A. Solomon, N. J. Majaj, and J. J. DiCarlo, "Deep Neural Networks Rival the Representation of Primate IT Cortex for Core Visual Object Recognition," PLoS Computational Biology, vol. 10, no. 12, 2014. [15] K. Simonyan and A. Zisserman, "Very deep convolutional networks for large-scale image recognition," arXiv preprint arXiv:1409.1556, 2014. [16] W. Ding, R. Wang, F. Mao, and G. Taylor, "Theano-based Large-Scale Visual Recognition with Multiple GPUs," in ICLR2015, 2015. [Online]. Available: http://arxiv.org/abs/1412.2302 [17] S. R. Lehky, R. Kiani, H. Esteky, and K. Tanaka, "Statistics of visual responses in primate inferotemporal cortex to object stimuli." Journal of Neurophysiology, vol. 106, no. 3, pp. 1097 -- 117, sep 2011. [18] N. K. Logothetis, J. Pauls, and T. Poggio, "Shape representation in the inferior temporal cortex of monkeys." Current Biology, vol. 5, no. 5, pp. 552 -- 563, 1995. [19] G. Y. Kov´acs, R. Vogels, and G. a. Orban, "Selectivity of macaque inferior temporal neurons for partially occluded shapes." The Journal of Neuroscience, vol. 15, no. March 1995, pp. 1984 -- 1997, 1995. [20] K. Tanaka, "Columns for complex visual object features in the infer- otemporal cortex: clustering of cells with similar but slightly different stimulus selectivities." Cerebral Cortex, vol. 13, pp. 90 -- 99, 2003. [21] M. Hasselmo, E. Rolls, G. Baylis, and V. Nalwa, "Object-centered encoding by face-selective neurons in the cortex in the superior temporal sulcus of the monkey," Experimental Brain Research, vol. 75, no. 2, pp. 417 -- 429, 1989. [22] W. Freiwald and D. Y. Tsao, "Functional compartmentalization and viewpoint generalization within the macaque face-processing system," Science, vol. 330, pp. 845 -- 851, 2010. [23] M. Ito, H. Tamura, I. Fujita, and K. Tanaka, "Size and position invariance of neuronal responses in monkey inferotemporal cortex." Journal of Neurophysiology, vol. 73, no. 1, pp. 218 -- 226, 1995. [24] H. Op De Beeck and R. Vogels, "Spatial sensitivity of macaque inferior temporal neurons," Journal of Comparative Neurology, vol. 426, pp. 505 -- 518, 2000. [25] K. J. Nielsen, N. K. Logothetis, and G. Rainer, "Dissociation between local field potentials and spiking activity in macaque inferior temporal cortex reveals diagnosticity-based encoding of complex objects," The Journal of neuroscience, vol. 26, no. 38, pp. 9639 -- 9645, 2006. [26] D. Zoccolan, D. D. Cox, and J. J. DiCarlo, "Multiple Object Response Normalization in Monkey Inferotemporal Cortex," Journal of Neuro- science, vol. 25, no. 36, pp. 8150 -- 8164, 2005. [27] E. Kobatake and K. Tanaka, "Neuronal selectivities to complex object features in the ventral visual pathway of the macaque cerebral cortex," Journal of neurophysiology, vol. 71, no. 3, pp. 856 -- 867, 1994. [28] B. Tripp, "A convolutional model of the primate middle temporal area," in ICANN, 2016. [29] D. B. Rubin, S. D. V. Hooser, and K. D. Miller, "The stabilized supralinear network: A unifying circuit motif underlying multi-input integration in sensory cortex," Neuron, vol. 85, no. 1, pp. 1 -- 51, 2015. [Online]. Available: http://dx.doi.org/10.1016/j.neuron.2014.12.026
1606.08370
1
1606
2016-06-27T17:16:25
The right time to learn: mechanisms and optimization of spaced learning
[ "q-bio.NC", "q-bio.MN" ]
For many types of learning, spaced training that involves repeated long inter-trial intervals (ITIs) leads to more robust memory formation than does massed training that involves short or no intervals. Several cognitive theories have been proposed to explain this superiority, but only recently has data begun to delineate the underlying cellular and molecular mechanisms of spaced training. We review these theories and data here. Computational models of the implicated signaling cascades have predicted that spaced training with irregular ITIs can enhance learning. This strategy of using models to predict optimal spaced training protocols, combined with pharmacotherapy, suggests novel ways to rescue impaired synaptic plasticity and learning.
q-bio.NC
q-bio
The right time to learn: mechanisms and optimization of spaced learning Paul Smolen, Yili Zhang and John H. Byrne Department of Neurobiology and Anatomy, W. M. Keck Center for the Neurobiology of Learning and Memory, The University of Texas Medical School at Houston, P.O. Box 20708, Houston, Texas 77030, USA Abstract For many types of learning, spaced training that involves repeated long inter-trial intervals (ITIs) leads to more robust memory formation than does massed training that involves short or no intervals. Several cognitive theories have been proposed to explain this superiority, but only recently has data begun to delineate the underlying cellular and molecular mechanisms of spaced training. We review these theories and data here. Computational models of the implicated signaling cascades have predicted that spaced training with irregular ITIs can enhance learning. This strategy of using models to predict optimal spaced training protocols, combined with pharmacotherapy, suggests novel ways to rescue impaired synaptic plasticity and learning. 1 Repetitive training helps form a long-term memory. Training or learning that includes long intervals between training sessions is termed spaced training or spaced learning. Such training has been known since the seminal work of Ebbinghaus to be superior to training that includes short inter-trial intervals (massed training or massed learning) in terms of its ability to promote memory formation. . Ebbinghaus stated1: "…with any considerable number of repetitions a suitable distribution of them over a space of time is decidedly more advantageous than the massing of them at a single time." His studies were based on the self-testing of acquired memory for lists of syllables, but the superiority of spaced training has now been established for many additional forms of human learning. For example, spaced learning is more effective than massed learning for facts, concepts and lists2, 3, 4, skill learning and motor learning5, 6, in classroom education (including science learning and vocabulary learning)7, 8, 9, and in generalization of conceptual knowledge in children10. Spaced training also leads to improved memory in invertebrates, such as the mollusk Aplysia californica 11, 12, 13, 14, Drosophila melanogaster15, 16 and bees17, and in rodents18, 19 and non-human primates20, 21. Memory extinction is commonly considered to involve the formation of a new memory, and in rat fear conditioning, spaced extinction trials are more effective than massed trails at establishing new memories22. Although it has been established that spaced training is superior to massed training in terms of inducing memory formation, key questions remain. What are the mechanisms underlying this superiority? Is it possible to use this mechanistic information to determine the optimal intervals between learning trials? If so, are fixed, expanding, or irregularly spaced intervals optimal? Another key question is whether an understanding of the mechanisms for optimal intervals can provide insights into the design of pharmacological approaches for memory enhancement. Computational models based on such a mechanistic understanding may be able to predict more complex approaches to memory improvement in which the application of multiple drugs, or combinations of drugs and training protocols, can enhance memory or treat deficits in learning and memory. In this Review, we describe how new insights from molecular studies may help explain the effectiveness of spaced training, and how the molecular findings relate to the traditional learning theories that aim to account for this effectiveness. We also review how models of signaling pathways that are involved in synaptic plasticity can suggest, and experiments 2 empirically validate, training protocols that improve learning and that rescue plasticity impaired by deficits of key molecular components. Finally, we discuss recent models that have suggested combined-drug therapies that may further enhance some forms of learning and that may have synergistic effects with optimized spaced learning on memory formation. Traditional learning theories We briefly summarize three of the well-known cognitive theories proposed to explain spaced training's superiority to massed training: encoding variability theory, study-phase retrieval theory and deficient-processing theory. Encoding variability theory23, 24, 25 posits that repeated stimulus presentations or learning trials are more likely to occur in multiple contexts if they are spaced further apart in time, and that a memory trace for repeated trials therefore includes elements of each of these contexts. Thus, spaced training would tend to bind together more contexts and hence form a more robust memory, as a greater number of testing contexts could elicit retrieval of the memory. Study-phase retrieval theory26, 27, 28, 29 posits that spaced stimulus presentations or learning trials are more effective than massed trials in reinforcing memory because each spaced trial elicits retrieval and reactivation of a memory trace formed by the preceding trial. In contrast, with short massed trials, the preceding memory trace is still active, so it is not retrieved or reactivated. Therefore memory cannot be reinforced. Study-phase retrieval theory also accounts for a decline in learning with excessively long intervals because in that case the preceding memory trace can no longer be retrieved. A recent variant, retrieved context theory, also incorporates elements of encoding variability theory and has succeeded in predicting the results of subsequently performed spaced learning experiments in humans30. Deficient processing theory posits that spaced training forms a stronger memory than does massed training because in the latter, some processes that are necessary to form memories are not effectively executed. The reasoning here becomes clearer by examining variants of this theory that specify the nature of the deficient process. One variant posits that excess habituation during massed trials prevents effective reinforcement of memory traces31, whereas others posit that there is a failure to consolidate a memory (also known as consolidation theory)32, 33, a lack of voluntary attention to massed presentations31, 34, or a lack 3 of cognitive rehearsals or reactivations of a memory trace within the short intervals that are characteristic of massed training27, 35. Consolidation occurs as a memory trace becomes more fixed and stable with time after training2, 36. Thus, consolidation theory37, 38, 39 posits that a long-term memory trace is more efficiently stabilized or strengthened by spaced trials. The lack of cognitive rehearsals variant of deficient-processing theory might also be considered a more specific form of consolidation theory, because it assumes that a minimum number of rehearsals, or autonomous reactivations, are required to consolidate a memory trace. Variants of deficient-processing theory and relevant experiments are further discussed in REFs28, 40, 41. Below, we focus substantially on consolidation theory because, of all the traditional learning theories, it seems to be most closely aligned with our current understanding of the cellular and molecular mechanisms of memory. Landauer37 was one of the first to develop a conceptual model of the ways in which consolidation principles could explain the effectiveness of spaced training. Although the model was originally developed to explain the effects of short spacing intervals on memory formation, it can readily be generalized for the effects of arbitrarily long intervals (FIG. 1). The model is based on two assumptions. First, the state of a neural circuit following the first learning trial is such that a second reinforcing trial soon after will not markedly increase the consolidation of the learning trace resulting from the first trial (FIG. 1a). Thus, in massed training, overlap between traces and, consequently, saturation of an unspecified molecular mechanism diminishes the traces' summed impact on the consolidation of memory. Only when the effects of the first trial decay can the effects of a second trial be fully expressed (FIG. 1b, 1c), leading to greater potential consolidation of a memory in spaced training than in massed training (greater net gain, FIG. 1d). The second assumption is that the probability that the second trial can successfully reinforce the first trial declines with time (FIG. 1e). Actual consolidation is the product of these two assumptions, yielding a prediction of an optimal interval for spaced learning (FIG. 1f). Peterson38 described a similar model focused on the dynamics of verbal learning. 4 Figure 1 Early conceptual model of how learning trace dynamics generate an optimal interval. a–c As described by the early model of Landauer37, spaced training is more effective at strengthening some form of trace corresponding to memory storage in the brain, although this conceptual model does not posit a biochemical or structural form for the trace. This model posits that memory formation becomes more effective with longer inter-stimulus intervals between training sessions because of decreasing temporal overlap between successive, short-lived learning traces. These learning traces do not themselves constitute a memory. However, their net effect, denoted here as net gain, contributes to the formation of a long-lived memory trace. In (a–c), learning traces elicited by two successive trials are shown. The model assumes that, for each value of the inter-trial interval length, a quantity denoted "net gain due to the reinforcing trial" is proportional to the red area. Shorter intervals are associated with more overlap of learning traces and less net gain. Therefore, a reinforcing trial is most effective after a refractory period following the preceding trial. For this conceptual theory, units for amplitude and time are arbitrary. d A greater summed effect, or net gain, of reinforcing trials occurs for longer inter-stimulus intervals. The effect reaches a plateau for long intervals as the overlap between successive learning traces goes to zero e Over longer times, a different quantity - the probability that a reinforcing trial will be effective at all in reactivating processes that constituted the preceding learning trace - declines. f An optimum interval for maximizing the strength of the long-lived memory trace results when the greater net gain of reinforcement at longer intervals (from d) is multiplied by the slowly declining probability that a reinforcement will reactivate a previous learning trace (from e). The optimum interval for net learning is the one that produces the peak level of the trace in f. Wickelgren39 extended consolidation theory by positing that the resistance of a memory trace to decay increases with the age of the trace over the total duration of a spaced learning protocol. Thus, a trace would become not only strong but also highly resistant to decay following spaced trials. Molecular traces of time Substantial progress has been made in understanding the molecular mechanisms of memory. Given this progress, in this section we focus on potential molecular mechanisms of the spacing effect on long-term memory formation. 5 There is now agreement that learning is implemented, at least in part, by changes in synaptic strength (synaptic plasticity). For example, fear-conditioned memories can be alternately erased and re-instantiated by long-term depression (LTD) and long-term potentiation (LTP), respectively, of a defined synaptic pathway42. Thus, the molecular processes that are essential for spaced learning might reinforce extant LTP. Reliable correlates of LTP are the remodeling and enlargement of postsynaptic dendritic spines, which are small protrusions that are associated with most excitatory synapses43. Thus, studying the differential dynamics of dendritic spine remodeling following massed versus spaced stimuli is likely to give insight into processes underlying the effectiveness of spaced training. Studies using rat hippocampal slice found that LTP induced by multiple trains of theta-burst stimuli was accompanied by extensive remodeling of synaptic ultrastructures44, 45 and that subsequent spaced trains of theta-burst stimuli with intervals of 60 min or more between trains, were needed for optimal reinforcement of LTP46. Stimulated dendritic spines were remodeled over > 1 h, leading to enlargement of the existing functional postsynaptic density45 and the presynaptic active zone44. The resulting increase in the numbers of AMPA-type and NMDA-type receptors at the synapse correlated with the magnitude of LTP. Two hypotheses that involve spine remodeling have been put forward to explain the greater efficacy of spaced trials over massed trials in memory formation. These hypotheses have a common theme, which is that the learning process includes a refractory period during which the second of two closely spaced stimuli would be ineffective in enhancing the effects of the first (FIG. 2a). One hypothesis is that spaced but not massed repetitions of a stimulus allow the refractory period to be overcome and lead to repeated enlargement of a set of spines and a strengthening of the synaptic connections mediated by these spines47. (FIG. 2b) A second, not mutually exclusive, hypothesis47, 48 is that molecular processes enable later spaced stimuli to induce LTP at spines that do not undergo initial enlargement. In this case spaced, but not massed, inter-trial intervals would allow for a molecular process termed 'priming' to be completed at these additional spines. After being primed, these spines would be strengthened by subsequent stimuli and incorporated into the memory trace (FIG. 2c). Currently, the molecular components of such a priming process are not known. 6 Figure 2 Model and hypotheses describing synaptic strengthening during spaced learning. a In the refractory-state model, spaced stimuli (left panel, stimulus 1, red arrow 1, followed substantially later by stimulus 2, red arrow 2) cumulatively strengthen a memory trace (as indicated by green line). By contrast, massed stimuli (right panel, stimulus 1 followed shortly after by stimulus 2) fail to cumulatively strengthen the memory trace. b The cumulative synaptic strengthening in spaced training may be due to progressive enhancement of long-term potentiation, which could result from successive increases in the in the strength of the same synaptic contacts (shown here as successive increases in the volume of the same postsynaptic dendritic spine). Thus in one of two current hypotheses describing synaptic strengthening during spaced learning, stimulus 1 enlarges a population of spines. If stimulus 2 follows shortly after the first stimulus (as in massed training), it cannot further affect spines. However, if stimulus 2 comes after a refractory period (as in spaced training), it can further enlarge the same population of spines. c Alternatively, enhancement of LTP could result from successive rounds of strengthening of new synaptic contacts. Thus in the second current hypothesis, stimulus 1 only enlarges a subset of affected spines, but primes additional spines. If stimulus 2 follows shortly after stimulus 1 (as in massed training), it has no effect. If stimulus 2 comes later (as in spaced training), it does not further enlarge the first subset of spines. Instead, stimulus 2 enlarges those spines that were primed, but not enlarged, by stimulus 1. Through use of Schaffer–commissural projections in rat hippocampal slices, two studies47, 48 have characterized the recruitment of additional synaptic contacts with the application of spaced stimuli. Theta burst stimuli applied at intervals of 10 min or 40 min did not 7 cumulatively increase LTP. However, for longer intervals (60 or 90 min), a cumulative increase in LTP was observed over three bursts of stimulation. Each theta burst stimulus (TBS) led to actin filament polymerization in spines, which is known to be important for the stabilization of LTP49. The second stimulus (TBS2) yielded polymerization in spines that were not apparently affected by the first stimulus (TBS1), if TBS2 followed TBS1 by 60 or 90 min. These data do not suggest that successive theta burst stimuli further strengthen the efficacy of the same spines. Instead, they suggest that TBS1 initiates priming at all synaptic contacts of the stimulated afferents but only initiates consolidation and strengthening at a subset of contacts. Spines that undergo priming but not consolidation exhibit a refractory period of ~ 60 min, suggesting that priming takes time to complete (FIG. 2a). If TBS2 is given after the refractory period, some or all of the primed spines undergo consolidation. These data are consistent with the second hypothesis presented in the preceding paragraph, because TBS1 appears to enlarge and strengthen some spines but at others, TBS1 only initiates priming. These primed spines can then be strengthened by TBS2. The dynamic properties of transcription factors and their interactions could also account for the superior efficacy of spaced training. LTP that persists for several hours or more requires translation and transcription50, 51, which is reliant on key transcription factors such as cAMP response element (CRE) binding protein (CREB)52. Spaced training may be more effective, in part, because it may allow transcription factors such as CREB sufficient time to be activated, bind to promoters and induce a round of transcription for the consolidation of LTP53 or long-term synaptic facilitation (LTF)54. In massed training, the trials would come too close together to initiate separate rounds of transcription. Indeed, in co-cultures of sensory and motor neurons from Aplysia, five spaced serotonin (5-HT) applications, each of 5 min with an inter-stimulus interval of 20 min (an analogue of spaced training) robustly elicits LTF that lasts for over 24 h14, whereas 5-HT applied continuously over 25 min (an analogue of massed training) fails to yield reliable LTF. In these sensory neurons, levels of a transcription activator CREB1 are elevated for at least 24 h after the spaced 5-HT treatments54, 55. This prolonged elevation of CREB1 owes to a positive feedback loop in which CREB1, acting via binding to a CRE regulatory element near creb1, increases the expression of creb154, 55 and other genes upregulated by CREB1. In addition, in these sensory neurons, the level of the transcription repressor CREB2 shows a late 8 drop at ~12 h after treatment56. This drop in CREB2, coupled with the rise in CREB1, plausibly corresponds to an increased potential for gene induction. Thus, an additional 5-HT pulse near 12 h after treatment might optimally reinforce LTF. LTF at these sensorimotor synapses is associated with a simple form of learning, long- term sensitization (LTS) of withdrawal reflexes. In vivo, four spaced electrical stimuli (30 min intervals between the stimuli) yield LTS that lasts > 24 h, with a weak residual of LTS being detectable at 4 d, and repetition of this spaced protocol once per day for 4 days yields much stronger LTS that lasts for over 1 week13, 57, 58. These data suggest that in this system, the dynamics of transcription activation and gene expression have slow components that can summate over multiple days, yielding long-lasting memory. Recent data also illustrates that in the hippocampus, CREB and CCAAT enhancer binding protein (C/EBP), another transcription factor that is important for LTP, can remain active for many hours after learning. Following inhibitory avoidance training in rats, late peaks in brain- derived neurotrophic factor (BDNF) expression and C/EBP expression occur at ~12 h post- training, and inhibiting BDNF action at this time blocks memory maintenance59. These BDNF dynamics result from a positive feedback loop in which C/ebp induction leads to Bdnf upregulation, with the resulting increase in BDNF further activating the C/EBP signaling pathway60. Although this slow feedback loop was activated by single-trial training rather than spaced training, it would be of interest to model these dynamics, and to examine whether an additional spaced trial at ~ 12 h post-training, leading to a second induction of C/EBP at the time of elevated C/EBP, might optimally reinforce learning. A second prediction would be that massed stimuli are less effective if repeated at an interval too brief to allow the transcription regulation and Bdnf expression necessary to activate this feedback loop. Insights that can be obtained from computational models of learning are discussed later in the article. time scale, On a shorter the dynamics of second messengers, kinases and phosphatases may contribute to the superiority of spaced training. One study with mice61 found marked phosphorylation and activation of CREB in the hippocampus and cortex when object recognition trials were separated by an interval of 15 min but not by an interval of 5 min. Protein phosphatase 1 (PP1) appeared to be necessary for this spacing effect, because PP1 inhibition allowed the shorter interval to activate CREB. A study involving Aplysia sensory neuron–motor neuron co-cultures62 found that protein kinase C (PKC) is activated to a greater extent during a 9 massed stimulus (continuous 5-HT application) than a spaced stimulus (15 min intervals between applications). It is known that PKC acts to downregulate protein kinase A (PKA) and that PKA activation is necessary for LTF, thus these data delineate crosstalk between signaling pathways such that LTF is suppressed, in part, by stronger PKC activation during massed training. Another study16 characterized the dynamics of MAP kinase (MAPK) and of MAPK phosphatase in Drosophila. In an olfactory learning protocol, each spaced training trial generated a distinct wave of MAPK activity, whereas massed training trials occurred too close together to generate distinct waves. The authors therefore hypothesized that effective learning depended on the generation of distinct waves, which was only seen after the spaced trials. Another phosphorylation-based mechanism has also been hypothesized to help explain the efficacy of spaced intervals in Drosophila. Spaced (15 min) intervals were more effective than massed (1 min) intervals in inducing olfactory learning, even given the same total training time (and thus more massed presentations)63. Two isoforms of Drosophila CREB - dCREB2-a and dCREB2-r - can activate and repress transcription, respectively. The authors proposed64 that the kinetics of the phosphorylation of these isoforms differed such that the kinase activation generated by less frequent, spaced trials was sufficient to phosphorylate and activate dCREB2- a, whereas dCREB2-r could only be effectively phosphorylated by massed trials. Thus, training involving spaced intervals could maximally activate transcription and possibly induce the formation of long-term memory by activating dCREB2-a but not the counteracting repressor dCREB2-r. Computational simulations have supported the plausibility of this mechanism65, but it has not been validated empirically. However, it appears likely that a similar type of mechanism that is based on competition between an activator of long-term memory formation and a repressor, with the repressor only activated at short intervals, might be needed to explain any similar data in which massed training is less effective than space training even given equal total training times. In experiments with Aplysia, when two electric shocks were given to induce LTS, maximal LTS was produced when the inter-stimulus interval was 45 min. LTS was not produced with intervals of 15 or 60 min66. The 45 min optimum was associated with activation of MAPK. Following either a single 5-HT pulse or a single shock, MAPK activation peaked at or near 45 10 min post-trial12, 66, thus a 45 min interval might optimally reinforce the effects of MAPK. It is known this delayed MAPK activation requires protein synthesis12, although the upstream mechanisms underlying the dynamics of the peak in MAPK activity at ~ 45 min are not well understood. Nevertheless, the key finding from these studies is that delayed activation of MAPK is intimately associated with the effectiveness of spaced stimuli to induce long-term memory. Similarly, training with intervals of 60 min, but not 20 or 120 min, enhanced object recognition learning in wild-type mice and in a mouse model of fragile X syndrome model (FMR1 knockout mice), at least partly by increasing synaptic activation of extracellular signal regulated kinase (ERK1 and ERK2: isoforms of MAPK)67. This 60 min interval was predicted to be optimal for learning because stimuli separated by 60 min had previously been found to enhance LTP in wild-type rodents47. Thus, in Aplysia, Drosophila, and mammals, MAPK activation appears to be a component of the molecular mechanism that underlies the spacing effect. Some of these molecular mechanisms appear to fit with a theory in which spaced training sessions are effective because they reinforce the same memory trace or group of strengthened synapses. However, spaced stimuli might also reinforce memory by recruiting new synapses. ERK1 and ERK2 (ERK1/2) activation is needed for some forms of LTP68 and one study69 compared ERK1/2 activation in rat hippocampal pyramidal neurons following three spaced tetanic bursts (at 5 min intervals) with that after three massed bursts (at 20 s intervals). About twice as many dendrites with active ERK1/2 were found following spaced bursts, suggesting that spaced trials may recruit additional synapses on different dendrites for LTP. Thus, a range of molecular and cellular mechanisms appears to contribute to the efficacy of spaced training, in parallel or in series. An extremely broad range of inter-trial intervals ranging from seconds to days has been used for spaced training (FIG. 3). For example, in honeybee olfactory learning, efficient spaced training can occur with intervals as short as 1 min17. Such brief intervals might allow for the reinforcement of the activity of a short-lived second messenger such as cAMP that is produced by preceding trials. The dynamics of kinase activation constitute a second substrate of spacing effects. In Aplysia, Drosophila, and mammals, the data discussed above indicate that commonly reported intervals, ranging from ~ 5 min to 1 h, may allow for the reinforcement of the activities of key kinases essential for LTP or LTF, and consolidate structural changes in dendritic spines. 11 in the relatively brief Figure 3 Different mechanisms may underlie enhancement of learning by intervals of widely varying spaced lengths. For inter-trial intervals (bottom trace), successive trials may coincide with and reinforce peak second messenger levels generated by preceding trials. In each trace, individual rectangles represent individual trials, and converging lines between traces represent the lengthening of time scales as one moves upwards illustration. For somewhat longer inter-trial intervals (several min to ~1 h), successive trials may reinforce the peak activities of kinases elicited by preceding trials and elicit LTP of primed dendritic spines. Intervals of this length may also, in hippocampus, be needed to allow replacement of inactivated receptors at stimulated spines82, enabling succeeding stimulus repetitions those spines. For intervals of ~ 1 h or more, succeeding trials may also align with peaks in transcription factor activity and gene expression owing to preceding trials. For the longest inter-trial intervals (many h or longer) succeeding trials may reactivate and thereby further potentiate consolidated memory traces. All these processes are likely to contribute to the consolidation of long-term memory, in many if not all species. However, depending on the inter-trial interval length used in a particular spaced learning protocol, the dynamics of a particular type of process (e.g. kinase activation) may contribute in particular to the efficacy of spaced learning. In addition, trials at one temporal domain (e.g., 1 day) may be unitary events, but also may constitute a block of spaced trials from another temporal domain (e.g., min to h). For example, an effective protocol for long-term sensitization training in Aplysia is the use of four trials with an intertrial interval of 30 min, with this block repeated four times with a one-day intertrial interval13. Thus, some effective training protocols consist of a hierarchy of temporal domains of training sessions, with briefer sessions embedded within longer ones. In this illustration intervals are shown with regular spacing, but more effective learning may occur with irregular spacing (see below). to potentiate It is plausible that the minimum inter-stimulus interval for effective learning, for a given protocol and system, corresponds to the interval that is necessary to allow each stimulus to contribute separately to a rate-limiting biochemical process. For example, for rapid honeybee olfactory learning with an effective interval of 1 min, the rate-limiting process might be second messenger accumulation or rapid activation of a kinase. For even shorter intervals, the time scale of the rate-limiting process might be too long to permit each brief stimulus to contribute separately to the process - a group of closely spaced stimuli would instead tend to act as just a single stimulus. For intervals of 1 min or more, each stimulus would be able to contribute a discrete increment to the rate-limiting process, allowing effective learning. For the spaced LTP protocol of Gall, Lynch and colleagues47, 48, an interval of 40–60 min is needed for successive TBS to further increase LTP. Here, the rate-limiting process would be different - plausibly slower activation of an unspecified kinase or other intracellular signaling event, with a time 12 constant near the minimum effective interval of ~ 40 min. Stimuli at intervals much shorter than this would not be able to generate summation of the rate-limiting process and would therefore not cause additional LTP. For other systems, a similar assumption may apply to the dynamics of transcription activation. For LTF and LTS in Aplysia, transcription, as discussed above, may constitute a rate- limiting process that helps determine the efficacy of spaced training. However, it is evident that even for systems such as honeybee olfactory learning that involve short, spaced intervals, effective long-term memory formation relies on the activation of transcription and translation, downstream of the intracellular signaling pathways that are activated by these intervals17, 70. Reactivation of memory traces may constitute an additional temporal substrate that underlies the longest reported effective intervals, on the order of a week71. Such intervals are likely to reactivate and reinforce consolidated patterns of strengthened synapses that correspond to memory traces that are maintained by neuronal network activity72. Spaced learning with these long intervals would reactivate critical components at these synapses, and in particular reactivate NMDA receptors at these synapses. Studies using inducible and reversible NMDA receptor knockouts have demonstrated that such NMDA receptor reactivation, which may also in part result from spontaneous neuronal activity, is required to sustain remote memory storage73, 74. Positive feedback loops that maintain key kinases and other molecules in persistently active states at strengthened synapses may also contribute to such long-term memory storage75, 76, 77, 78, 79. An important topic for future research will be to further investigate the molecular processes that support effective spaced learning for humans that involves inter-trial intervals of a day or more. An implication of Fig. 3 is that multiple temporal domains of spaced training can be engaged in spaced training. For example, an effective protocol for long-term sensitization training in Aplysia is the use of four trials with an inter-trial interval of 30 min, repeated four times with a one-day inter-trial interval13. Thus, at least in some cases, there appears to be a hierarchy of temporal domains of training protocols with briefer protocols embedded within longer ones. The above considerations, and most empirical studies, are concerned with only typical, or minimum, inter-trial or inter-stimulus intervals for effective spaced learning or for the summation of LTP. Only a few studies have delineated, for any specific system (that is, a given species and stimulus protocol), both minimum and maximum effective intervals. One study80 13 found that in a hippocampal slice preparation, 5–10 min intervals between tetani were ideal for induction of LTP and they produced similar levels of LTP, with longer or shorter intervals yielding both less LTP and less ERK1/2 activation. In Aplysia, LTS was effectively induced by an interval of 45 min between electrical stimuli, but not by intervals of 15 min or 60 min66. As noted above, the authors of this study hypothesized that the coincidence of peak MAPK activation with the second trial was necessary for effective learning. In addition, 60 min intervals were effective for forming object location memory in mice with three trials, but intervals of 20 min or 120 min were not67. Owing to the small number of such studies and the lack of sufficient characterization of the accompanying molecular processes, it is not yet possible to make detailed statements about the ways in which intracellular signaling pathways could co-operate to generate both minimum and maximum intervals. For maximum intervals, a reasonable qualitative assumption is that each trial or stimulus generates a separate, relatively short-lived biochemical trace, and that for effective spaced learning, these traces must overlap and summate, with the summed magnitude driving the formation of long-lasting synaptic potentiation. These dynamics would be analogous to the necessary overlap of traces in the conceptual model of Landauer (FIG. 1a-c). For intervals longer than the maximum, the individual biochemical traces would decay and not overlap. Recent data and learning theories Do the biochemical and morphological mechanisms proposed to contribute to the greater efficacy of spaced training align with traditional cognitive theories? At this point, much of the extant cellular data seem to be compatible with the deficient processing theory, particularly two of its variants: the consolidation theory and the lack of cognitive rehearsals theory. In the consolidation theory, intervals between massed trials are proposed to be too short for the consolidation and consequent summation of memory traces that are engendered by successive trials. In the cognitive rehearsals theory, massed trials are proposed to lead to fewer cognitive rehearsals, or autonomous reactivations, than do spaced trials, and therefore less cumulative consolidation and persistence of a memory. The required refractory period of ~ 1 h between successive theta-burst stimuli in to induce progressive increments in hippocampal LTP46, 47 may be in line with the first of these variants, that short intervals are insufficient for consolidation and consequent summation of 14 memory traces. The refractory period appears to be necessary to complete the 'priming' of dendritic spines that were stimulated, but not potentiated, by the first theta-burst. Priming allows these spines to potentiate after the second burst, and thus constitutes a biochemical stimulus trace (FIG. 1d and FIG. 2a). Kramár et al.47 noted that in hippocampal slices, additional potentiation can be induced up to four hours after induction of the first LTP increment81. The stimulus trace associated with priming may therefore take at least four hours to decay. Such a long trace lifetime might allow a broad temporal window for optimal training trials. In rat hippocampal slices, theta burst stimuli lead to proteolytic inactivation of integrin receptors at stimulated dendritic spines82. These receptors are then replaced by vesicular transport of new receptors over ~ 40 – 60 min, and it is hypothesized82 that subsequent theta burst stimuli at these synaptic contacts cannot induce spine enlargement or LTP until after this replacement has occurred, thus accounting for the refractory period of ~ 1 h in order for a second theta burst stimulus to yield additional LTP. This receptor replacement may constitute, at least in part, the priming of dendritic spines discussed above. These hypothesized dynamics may be in line with deficient processing theory, with receptor replacement being the necessary process that can only occur during spaced inter-trial intervals (FIG. 3). Transcription factor activation also constitutes a biochemical trace, and in some systems, training may only be effective if inter-trial intervals are long enough so that each trial can induce a separate round of transcription and translation. Similarly, short (massed) inter-trial intervals may not lead to sufficient levels, or a sufficient duration, of activated MAPK or other kinases to support the consolidation of long-term memory. The variant of deficient processing theory positing that only spaced trials can generate sufficient cognitive rehearsals or reactivations of a memory to support long-term memory consolidation may also correspond with the empirical finding that repeated theta-burst stimuli, spaced by ~ 1 h, can recruit additional dendritic spines by potentiating spines that were primed by preceding stimuli. A memory reactivation would be analogous to theta-burst stimulus in that both events would initiate priming and potentiation. It also appears plausible that repeated memory reactivations might induce further rounds of transcription of genes involved in LTP, such as C/ebp and other CREB-activated genes, supporting further consolidation of long-term memory. 15 To more strongly connect this variant of deficient-processing theory to recent cellular and molecular data, one must also assume that reactivations of a memory reactivate some of the same neurons and synapses that were activated in the original learning sessions. In that way, the rehearsals and learning trials would reinforce memory in the same way. This assumption appears plausible but requires further empirical investigation. Although finer-grain analyses are necessary, a study using functional MRI during verbal learning supports this assumption83. In this study, a specific brain region associated with rehearsal of verbal memory, the left frontal operculum, was activated more during spaced learning of paired-word associations than during massed learning. We note that these posited memory reactivations, on time scales of ~ 1 h or longer, are distinct from voluntary rehearsals of a memory on a short time scale (seconds or ~ 1 min). Substantial behavioral evidence suggests that this latter voluntary, short-term rehearsal is not essential for spaced learning31, 34. The remaining variants of deficient processing theory, which focus on habituation or on a lack of voluntary attention during massed presentations, do not appear to relate as readily to the current single-neuron data. These variants have also been argued to not readily accommodate certain verbal learning observations2. As regards encoding-variability theory, data on neuronal network dynamics, rather than single-neuron data, will be needed to determine to what extent the binding of contexts to memory occurs, which is required in this theory. Similar data will also be needed to assess whether the binding of memories of later trials to those of earlier trials occurs, which is required in study-phase retrieval theory. It will be important to re-assess all these competing spaced learning theories as more information becomes available on the dynamics of memory networks. Indeed, different theories may be more or less applicable to different memory systems. Irregular spacing can enhance learning Attempts to optimize the spacing effect have generally been based on trial-and-error approaches. Consequently, most, if not all, training protocols used in animal and human studies are probably not optimal. For almost all learning paradigms, the training intervals are fixed, although in one type of spaced training paradigm, the intervals between sessions progressively lengthen2, 84. However, a meta-analysis2 and a text learning study84 found no substantial evidence for the superiority of this approach in terms of promoting long-term memory formation. 16 It appears evident that at least part of the improvement in learning that is found with spaced training protocols can be explained by the dynamic relationships between the training trials and the underlying cellular and molecular mechanisms that are associated with memory formation (FIG. 3). But is the inverse possible? Can knowledge of the dynamics of the memory mechanisms be used to enhance memory processing by predicting optimal training protocols, possibly with irregular training intervals? One approach is to develop models of the biochemical cascades that underlie memory formation and use simulations to rapidly test the effectiveness of different training protocols85. In recent years, models have described the dynamics of the biochemical reactions that transduce stimuli into LTP86, 87, 88. These models have differential equations that simulate and predict the dynamics of the activities of key molecular species. Simulations have reproduced the dynamics of MAPK during LTP induction80, 86, 88. Models have also simulated the activity time courses of PKA, CAMKII, other key enzymes and downstream transcription factors during LTP induction88, 89, 90. Each signaling cascade in these models displays a characteristic activity time course, thus it is likely some irregular sequence of intervals would be predicted to maximize the induction of LTP. For example, subsequent trials that are delivered at times that coincide with kinase activity peaks might optimally reinforce learning. One study from our laboratory developed a model describing the 5-HT-induced PKA and ERK signaling pathways essential for LTF in Aplysia85. In the model (FIG. 4a), the necessity of PKA and ERK activation for LTF was simply represented with a variable termed 'inducer'. The value of inducer was proportional to the product of PKA and MAPK activities. The amount of LTF and LTS was predicted to increase with an increase in the peak value of inducer. Ten thousand different protocols consisting of five trials that were separated by intervals of 0 to 45 min were simulated (FIG. 4b). The ability to simulate and predict the effects of so many protocols in a relatively short space of time represents a distinct advantage of computational studies over empirical studies. The simulations determined that of these protocols, a massed protocol (FIG. 4b) produced the lowest peak value of inducer, consistent with data that massed 5-HT application fails to produce LTF14. The 'best' protocol yielding the highest peak value for inducer, termed the 'enhanced' protocol (FIG. 4b), had irregular intervals. The protocol termed the 'standard' protocol (five 5-min pulses of 5-HT, with uniform inter-stimulus intervals of 20 min) (FIG. 4b), has been commonly used to induce LTF in empirical studies for ~30 years91. This standard protocol yielded an intermediate peak value of inducer and was predicted to have an 17 intermediate effectiveness (FIG. 4c). These predictions were empirically validated. LTF and LTS produced by the enhanced protocol exceeded that from the standard protocol (FIG. 4d)85. An explanation for this enhancement of LTF, consistent with data11, 12, 92, is as follows. In response to each 5-HT pulse, PKA activates rapidly and decays rapidly (FIG. 4c). MAPK activity rises and decays more slowly, not peaking until ~45 min after a pulse. The four initial pulses initiate a surge of MAPK activity, which peaks near the time of the last pulse. This last pulse activates PKA, so that a PKA peak is approximately coincident with peak MAPK activation, maximizing inducer and the predicted LTF. Figure 4 Dynamics of a model which has successfully predicted greater efficacy for a learning protocol with irregular spaced intervals. a A simplified mathematical model85 describes the activation and effects of two key kinases necessary for LTF, a cellular correlate of a simple form of learning, long- term sensitization. Brief applications of 5-HT activate protein kinase A (PKA) via elevation of the secondary messenger cAMP, and activate the ERK isoform of MAP kinase via a Ras-Raf-MEK cascade. PKA and ERK interact, at least in part via phosphorylation of transcription factors, to induce LTF. In the model, the variable 'inducer' represents the PKA–ERK interaction. A higher peak value of inducer was assumed to predict a greater amplitude of LTF. b Six samples of the 10,000 5-HT protocols that were simulated with the model. All protocols consist of five 5-min pulses of 5-HT, diagrammed as square 18 waves, with inter-pulse intervals chosen as multiples of 5 min, in the range from 5–50 min. The standard protocol (green trace) is the protocol most commonly used in studies of LTF in vitro. The enhanced protocol (red trace) produced the largest peak value of inducer, whereas the massed protocol (blue trace) produced the smallest peak value of inducer. The standard protocol has uniform inter-pulse intervals of 20 min, whereas the enhanced protocol has non-uniform intervals of 10, 10, 5 and 30 min. The massed protocol has no gaps between the 5-HT pulses. c Simulated time courses of activated PKA, activated ERK and inducer in response to the standard protocol (green traces), enhanced protocol (red traces), and massed protocol (blue traces). The unit of concentration (y axes) is µM. d In an empirical validation of the model's prediction, the LTF induced by the enhanced protocol, as determined by the percentage increase in the amplitude of excitatory postsynaptic potentials (EPSPs), was greater than the LTF produced by the standard protocol. Figure adapted from REF. 85. for several factors transcription If irregularly spaced protocols can enhance normal learning, might modeling also predict protocols capable of restoring learning that is impaired by a genetic mutation or other physiological insults? A recent study90 tested this hypothesis. CREB binding protein (CBP) is an acetyltransferase and essential co-activator including phosphorylated CREB (pCREB). CBP is also required for the consolidation of long-term memory93. Mutations that decrease CBP activity cause a human genetic disorder termed Rubinstein–Taybi syndrome (RTS)94 that is associated with intellectual disability and learning deficits, and Cbp+/− mice show impaired LTP and LTM95. The recent study90 used small interfering (siRNA) knockdown of CBP in Aplysia sensory neurons to impair LTF. In this study, the model previously used85 to predict optimal, irregularly spaced protocols that would enhance LTF (FIG. 4a) was extended to represent induction of c/ebp, a transcription factor known to be essential for LTF96. In simulations of the effects of different spaced protocols, greater peak levels of phosphorylated C/EBP (pC/EBP) were taken to predict greater LTF. Simulations showed a substantial decrease in pC/EBP when CBP was reduced by a decrement that corresponded to the siRNA effect. A 'rescue' protocol with irregularly spaced intervals was predicted to restore peak pC/EBP and, correspondingly, LTF. This rescue protocol was empirically validated to restore normal LTF in Aplysia. A similar predicted rescue protocol of irregularly spaced intervals rescued a deficit in LTF that was produced by a siRNA knockdown of CREB197. Although these empirical studies were conducted in Aplysia, it should be noted that key molecular mechanisms of memory are substantially conserved from simple model organisms such as Aplysia to mammals52, 96. For example, LTF and LTP both rely on PKA and ERK activation12, 92, 98 and both rely on co-operative gene induction by pCREB and CBP53, 96, 99, 100. LTF relies on deactivation of CREB2, a transcriptional repressor101. Similarly, relief of 19 transcriptional repression due to ATF4, a mammalian analogue of CREB2, appears to be important for the maintenance of hippocampal LTP102, 103. Thus, the results with Aplysia suggest it may be possible, in complex organisms including mammals, to computationally predict the efficacies of numerous learning or training protocols, a process that is impractical using empirical studies alone. Given that knowledge of the underlying biochemical cascades can help develop models to predict optimal training protocols, can models also be used to predict pharmacological targets to improve memory? The time may also be right for such an approach. For example, if simulated LTP deficits were rescued by combined parameter changes corresponding to known drug effects, these 'best' parameter combinations might prioritize drug combinations for testing in animal models. A recent study104 took a first step by modelling LTP induction and transcriptional regulation by CREB, and simulating the effects of drugs on LTP by altering the model's parameters. In this model, the magnitude of LTP induction was represented by an increase in a synaptic weight variable. LTP impairment seen in a mouse model of RTS95 was first simulated. Then, starting from this simulation, the parameters were altered in ways corresponding to plausible single-drug effects. However, no single-drug effect completely rescued LTP. Therefore, pairs of parameter changes were considered, corresponding to plausible paired-drug effects. Two pairs were identified that restored LTP. In the first case, an increased rate constant to application of an acetyltransferase activator, was paired with an increased duration of stimulus-induced cAMP elevation, corresponding to application of a cAMP phosphodiesterase (PDE) inhibitor. The second pair corresponded to a PDE inhibitor paired with a deacetylase inhibitor. For both pairs, additive synergism, defined as a combined-drug effect that exceeds the summed effects of the individual drugs, was also evident, as quantified by a simple additive measure (FIG. 5). A subsequent empirical study by another group did find that pairing a PDE inhibitor with a deacetylase inhibitor was effective in rescuing a deficit of LTP in a mouse model of Alzheimer disease105. A further extension of these strategies might similarly predict, and empirically test, enhancement of synaptic plasticity when pharmacotherapy is combined with computationally designed spaced protocols. for histone acetylation, corresponding Figure 5 A model predicts that a pair of drugs can act synergistically to enhance LTP. A CREB binding protein (Cbp) mutation impairs hippocampal LTP and impairs learning in mice, and Cbp+/- mice are considered a model for aspects of Rubinstein- 20 Taybi syndrome in humans104. We developed a model to examine whether drugs could be used to overcome this impairment in LTP. This figure was generated from a series of simulations of the effects of two drugs on the induction of LTP. LTP was modeled as the percent increase in a synaptic weight variable. In the absence of drugs, simulated LTP induced by a high-frequency tetanic stimulus was strongly impaired. Only a 50% increase in synaptic weight for Cbp+/- occurred, compared to an increase in synaptic weight of 148% with non-mutated Cbp. The effect of each drug was simply modeled as a change in the value of a kinetic parameter. In this series of simulations, the doses of two drugs - drug 1, a cAMP phosphodiesterase inhibitor, and drug 2, an acetyltransferase activator - were concurrently varied. The effect of drug 1 was simulated by decreasing a rate constant for cAMP degradation, and the effect of drug 2 was simulated by increasing a rate constant for histone acetylation. The 'dose' of drug 1 - the amplitude of the rate constant change - was increased and simultaneously the dose of drug 2 was decreased. 80 pairs of drug doses were simulated. Both drugs substantially enhanced LTP. For drug 2 alone (left endpoint of graph) LTP was 155%, and for drug 1 alone (right endpoint), LTP was 116%. For both drugs together, with smaller doses of each drug, intermediate LTP amplitudes were observed (combined effect curve). This series of simulations further shows that additive synergism persists over a substantial range of drug doses. Additive synergism is quantified as the difference (red arrow) between the LTP simulated when both drugs are applied together (combined effect curve), and the LTP simulated by adding together the effect of the drugs applied individually in separate simulations (summed effects curve). Future directions There is reason for optimism that more predictive models for determining optimal intervals between learning trials will be available in the near future, because the molecular data that are necessary for the development of such models, delineating dynamics of signaling pathways that are important for LTP and long-term memory, continue to accumulate rapidly. However, despite the progress being made in understanding the molecular mechanisms of the spacing effect, some aspects of this effect cannot be explained by current models and constitute important directions for future research. For example, in human verbal learning, an interesting positive correlation exists between the length of inter-trial intervals for effective spaced learning and the retention interval (that is, the interval between the final training trial and the test of memory retention). With relatively short retention intervals (~1 min–2 h), training intervals in the broad range of ~1 min to 3 h yield greater verbal learning than do training intervals of 2 d or more2. With a longer retention interval of 1 d, a 1 d training interval yielded greater learning than did a very short (< 30 s) interval. For verbal learning with a retention interval of 6 months, a training interval of 7 d was superior to an interval of 3 d71. This correlation between longer training and retention intervals suggests that longer training intervals preferentially form a memory trace with a very long lifetime. For the temporal range of minutes versus hours, it is plausible that a longer trace lifetime corresponds, at least in part, to increased activation of transcription by the longer training intervals. However, this explanation may not suffice when comparing training intervals of ~ 1 d versus many days. It would be of interest to determine whether reactivation of stored 21 memory representations at the network level, or transfer of these representations between brain regions, contribute to this correlation. Another challenge will be to use innovative strategies to test the predictions of the cognitive theories for the spacing effect. For example, consider the variant of deficient- processing theory positing that repeated cognitive rehearsals of a memory are needed for consolidation. A neuronal correlate of rehearsals is, plausibly, repeated activation of a specific neuron assembly that serves as a locus of storage of a long-term memory trace. Empirically, is such repeated activation necessary for persistence of memory for days or longer? Repeated spontaneous activation of neuron assemblies does occur106, 107 as does repeated replay or rehearsal of assemblies that encode recent experiences108, 109. One study supporting the necessity of such replay found that the post-training suppression of activity of neurons that were engineered to overexpress CREB in the amygdala for several h after conditioning blocked the consolidation of a memory of association between cocaine and a location110. Similar blocking effects were obtained by the indiscriminate activation of neurons that overexpressed CREB. Although encouraging, these manipulations lack the cellular precision that is necessary to demonstrate conclusively that reactivation of a particular assembly of neurons is essential for the persistence of long-term memory. Future studies using optogenetic techniques could provide that precision. Similarly, innovative strategies will be needed to address whether effective spaced learning requires the binding of contextual and episodic memories at the neuronal network level, such as posited by encoding variability theory, or increased binding due to greater retrieval effort, as posited by study-phase retrieval theory. The successful prediction of the interval structure of behavioral training protocols that may overcome some human learning deficits (when applied alone or in combination with pharmacotherapy) will require improved knowledge of the signaling pathways that underlie LTP and long-term memory and of the ways in which the deficits affect those pathways. Future models are still likely to be incomplete owing to gaps in knowledge. For example, data will be incomplete and associated with unavoidable uncertainties in the values of biochemical parameters in models such as enzyme activities or protein concentrations. In model development, data from several preparation types (for example, cell cultures and slices) and species (for example, primates and rodents) commonly need to be used to estimate different parameters86, 88. However, although these limitations are important, the potential benefits of combining modeling with experiments in the ways discussed herein are extensive, such that this 22 strategy may have promise for improving the clinical and educational outcomes for patients with learning and memory deficits. In addition, it is possible that education and learning in individuals without such deficits could benefit from such a strategy. Indeed, enhancing normal learning by judicious pharmacotherapy has recently received attention111, and combining drugs with optimized spaced learning protocols might yield even better outcomes. Glossary Reinforcement: A broad term used here to describe a stimulus or item that enhances the strength or lifetime of a memory. Reinforcement stimuli activate biochemical and molecular processes and thereby regulate changes in synaptic strength associated with memory. Memory Extinction: The decline of a learned behavioral response to a conditioned stimulus following the withdrawal of reinforcement stimuli that were previously paired with repetitions of the conditioned stimulus. This term is distinct from forgetting, which is the decline of a learned response during the prolonged absence of stimulus repetitions. Habituation: A decrease of the behavioral response to a stimulus following frequent repetitions of that stimulus. This term is distinct from extinction, because habituation can denote a decrease of response to a stimulus that was never paired with a reinforcing stimulus. Memory Reactivation: Re-instantiation of a conditioned behavioral response or of neural activity associated with a specific response. Reactivation can be elicited by presentation of a conditioning stimulus or of the context in which learning previously occurred, or it can be spontaneous, occurring as a part of normal ongoing neural activity. Drug Synergism: In combined-drug treatment, a synergistic effect of the combination is an effect greater than what would be predicted by considering the individual drugs as independent and not interacting. Several methods for assessing drug synergism exist. The method referred to here (REF. 104) is one of the simplest. References 23 1. Ebbinghaus, H. Memory: a Contribution to Experimental Psychology, Chapter 8, p. 89 (Teachers College, Columbia University, New York, 1885/1913) (reprinted Thoemmes Press, Bristol, 1999). 2. Cepeda, N. J., Pashler, H., Vul, E., Wixted, J. T. & Rohrer, D. Distributed practice in verbal recall tasks: A review and quantitative synthesis. Psychol. Bull. 132, 354-380 (2006). This extensive review and meta-analysis delineates the comprehensive body of knowledge describing human spaced and massed verbal learning, as well as theories posited to explain spaced training's superiority to massed training in terms of inducing long-term memory formation. 3. Godbole, N. R., Delaney, P. F. & Verkoeijen, P. P. The spacing effect in immediate and delayed free recall. Memory 22, 462-469 (2014). 4. Raman, M., McLaughlin, K., Violato, C., Rostom, A., Allard, J. P. & Coderre, S. Teaching in small portions dispersed over time enhances long-term knowledge retention. Medical Teacher 32, 250-255 (2010). 5. Donovan, J. J. & Radosevich, D. J. A meta-analytic review of the distribution of practice effect. J. Appl. Psychol. 84, 795-805 (1999). 6. Shea, C. H., Lai, Q., Black, C. & Park, J. H. Spacing practice sessions across days benefits the learning of motor skills. Human Movement Sci. 19, 737-760 (2000). 7. Gluckman, M., Vlach, H. A. & Sandhofer, C. M. Spacing simultaneously promotes multiple forms of learning in children's science curriculum. Appl. Cognit. Psychol. 28, 266-273 (2014). 8. McDaniel, M. A. & Callender, A. A. Cognition, memory, and education. In: Learning and Memory: A Comprehensive Reference, Vol. 2 (ed. Roediger, H. L.) 819-844 (Academic Press, Oxford, 2008). 9. Sobel, H. S., Cepeda, N. J. & Kapler, I. V. Spacing effects in real-world classroom vocabulary learning. Appl. Cognit. Psychol. 25, 763-767 (2011). 10. Vlach, H. A. The spacing effect in children's generalization of knowledge: allowing children time to forget promotes their ability to learn. Child Devel. Persp. 8, 163-168 (2014). This study found that, in children's education, spacing of learning sessions promoted not only better fact retention but also generalization of currently learned items to bind with and extend previously established conceptual categories. 24 11. Philips, G. T., Kopec, A. M, & Carew, T. J. Pattern and predictability in memory formation: from molecular mechanisms to clinical relevance. Neurobiol. Learn. Mem. 105, 117-124 (2013a). 12. Philips, G. T., Ye, X., Kopec, A. M. & Carew, T. J. MAPK establishes a molecular context that defines effective training patterns for long-term memory formation. J. Neurosci. 33, 7565- 7573 (2013b). This work characterized the dynamics of MAPK and its activation of C/EBP, a transcription factor that is important for LTM, identifying the response of MAPK within a specific time window as essential for generating an optimal inter-trial interval. 13. Wainwright, M. L., Zhang, H., Byrne, J. H. & Cleary, L. J. Localized neuronal outgrowth induced by long-term sensitization training in Aplysia. J. Neurosci. 22, 4132-4141 (2002). 14. Mauelshagen, J., Sherff, C. M. & Carew, T. J. Differential induction of long-term synaptic facilitation by spaced and massed applications of serotonin at sensory neuron synapses of Aplysia californica. Learn. Mem. 5, 246–256 (1998). 15. Beck, C. D., Schroeder, B. & Davis, R. L. Learning performance of normal and mutant Drosophila after repeated conditioning trials with discrete stimuli. J. Neurosci. 20, 2944- 2953 (2000). 16. Pagani, M. R., Oishi, K., Gelb, B. D. & Zhong, Y. The phosphatase SHP2 regulates the spacing effect for long-term memory induction. Cell 139, 186-198 (2009). 17. Menzel, R., Manz, G., Menzel, R. & Greggers, U. Massed and spaced learning in honeybees: the role of CS, US, the intertrial interval, and the test interval. Learn. Mem. 8, 198-208 (2001). 18. Anderson, M. J., Jablonski, S. A. & Klimas, D. B. Spaced initial stimulus familiarization enhances novelty preference in Long-Evans rats. Behav. Processes 78, 481-486 (2008). 19. Bello-Medina, P. C., Sanchez-Carrasco, L., Gonzalez-Ornelas, NR., Jeffery, K. J. & Ramirez-Amaya, V. Differential effects of spaced vs. massed training in long-term object- identity and object-location recognition memory. Behav. Brain Res. 250, 102-113 (2013). 20. Medin, D. L. The comparative study of memory. J. Human Evol. 3, 455-463 (1974). 21. Robbins, O. & Bush, C. T. Memory in great apes. J. Exp. Psychol. 97, 344-348 (1973). 25 22. Urcelay, G. P., Wheeler, D. S. & Miller, R. R. Spacing extinction trials alleviates renewal and spontaneous recovery. Learn. Behav. 37, 60-73 (2009). 23. Benjamin, A. S. & Tullis, J. What makes distributed practice effective? Cogn. Psychol. 61, 228-247 (2010). 24. Madigan, S. A. Intraserial repetition and coding processes in free recall. J. Verb. Learn. Verb. Behav. 8, 828-835 (1969). 25. Melton, A. W. The situation with respect to the spacing of repetitions and memory. J. Verb. Learn. Verb. Behav. 9, 596-606 (1970). 26. Braun, K. & Rubin, D. C. The spacing effect depends on an encoding deficit, retrieval, and time in working memory: evidence from once-presented words. Memory 6, 37-65 (1998). 27. Greene, R. L. Spacing effects in memory: Evidence for a two-process account. J. Exp. Psychol. Learn. Mem. Cogn. 15, 371-377 (1989). 28. Hintzman, D. L., Summers, J. J. & Block, R. A. Spacing judgments as an index of study- phase retrieval. J. Exp. Psychol. Hum. Learn. 1, 31-40 (1975). 29. Tzeng, O. J. & Cotton, B. A study-phase retrieval model of temporal coding. J. Exp. Psychol. Hum. Learn. 6, 705-716 (1980). 30. Siegel, L. L. & Kahana, M. J. A retrieved context account of spacing and repetition effects in free recall. J. Exp. Psychol. Learn. Mem. Cogn. 40, 755-764 (2014). 31. Hintzman, D. L. Theoretical implications of the spacing effect. In: Theories in Cognitive Psychology: The Loyola Symposium (ed. Solso, R. L.) 77-90 (Erlbaum, Potomac, 1974). 32. Dempster, F. N. Spacing effects and their implications for theory and practice. Educ. Psychol. Rev. 1, 309-330 (1989). 33. Toppino, T. C. The spacing effect in young children's free recall: support for automatic- process explanations. Mem. Cogn. 19, 159-167 (1991). 34. Hintzman, D. L. Repetition and memory. In: The Psychology of Learning and Motivation, Vol. 10 (ed. Bower, G. H.) 47-91 (Academic Press, New York, 1976). 35. Rundus, D. Analysis of rehearsal processes in free recall. J. Exp. Psychol. 89, 63-77 (1971). 36. Lechner, H. A., Squire, L. R. & Byrne, J. H. 100 years of consolidation-remembering Müller and Pilzecker. Learn. Mem. 6, 77-87 (1999). 26 37. Landauer, T. K. Reinforcement as consolidation. Psychol. Rev. 76, 82-96 (1969). This early paper posited and described elements of the dynamics of spaced learning, such as the saturation of memory formation by closely spaced trials and the failure of reinforcement for long trial spacing, and how these elements generate an optimal inter-trial interval. 38. Peterson, L. R. Short-term verbal memory and learning. Psychol. Rev. 73, 193-207 (1966). 39. Wickelgren, W. A. Trace resistance and the decay of long-term memory. J. Math. Psychol. 9, 418-455 (1972). 40. Cuddy, L. J. & Jacoby, L. J. When forgetting helps memory: an analysis of repetition effects. J. Verb. Learn. Verb. Behav. 21, 451-467 (1982). 41. Toppino, T. C. & Bloom, L. C. The spacing effect, free recall, and two-process theory: a closer look. J. Exp. Psychol. Learn. Mem. Cogn. 28, 437-444 (2002). 42. Nabavi, S., Fox, R., Proulx, C. D., Lin, J.Y., Tsien, R. Y. & Malinow, R. Engineering a memory with LTD and LTP. Nature 511, 348-352 (2014). 43. Bosch, M. & Hayashi, Y. Structural plasticity of dendritic spines. Curr. Opin. Neurobiol. 22, 383-388 (2012). 44. Bell, M. E., Bourne, J. N., Chirillo, M. A., Mendenhell, J. M., Kuwajima, M. & Harris K. M. Dynamics of nascent and active zone ultrastructure as synapses enlarge during LTP in mature hippocampus. J. Comp. Neurol. 522, 3861-3864 (2014). This work, one of a group of related studies by the Harris laboratory, characterized in detail pre- and postsynaptic structural changes that are necessary for the consolidation of LTP. 45. Bourne, J. N. & Harris, K. M. Coordination of size and number of excitatory and inhibitory synapses results in a balanced structural plasticity along mature hippocampal CA1 dendrites during LTP. Hippocampus 21, 354-373 (2011). 46. Cao, G. & Harris K. M. Augmenting saturated LTP by broadly spaced episodes of theta-burst stimulation in hippocampal area CA1 of adult rats and mice. J. Neurophysiol. 112, 1916- 1924 (2014). 27 47. Kramár, E. A., Babayan, A. H., Gavin, C. F., Cox, C. D., Jafari, M., Gall, C. M., Rumbaugh, G. & Lynch, G. Synaptic evidence for the efficacy of spaced learning. Proc. Natl. Acad. Sci. USA 109, 5121-5126 (2012). This study, together with REF. 48, substantiates the key hypothesis that spaced stimuli first 'prime' and then strengthen upon repetition successive groups of dendritic spines; creating a morphological correlate of the memory trace that spaced, but not massed, protocols effectively consolidate. 48. Lynch, G., Kramár, E. A., Babayan, A. H., Rumbaugh, G. & Gall, C. M. Differences between synaptic plasticity thresholds result in new timing rules for maximizing long-term potentiation. Neuropharmacology 64, 27-38 (2013). 49. Rex, C. S., Chen, L. Y., Sharma, A., Liu, J., Babayan, A. H., Gall, C. M. & Lynch, G. Different Rho GTPase-dependent signaling pathways initiate sequential steps in the consolidation of long-term potentiation. J. Cell Biol. 186, 85-97 (2009). 50. Nguyen, P. V., Abel, T. & Kandel, E. R. Requirement of a critical period of transcription for induction of a late phase of LTP. Science 265, 1104-1107 (1994). 51. Scharf, M. T., Woo, N. H., Lattal, K. M., Young, J. Z., Nguyen, P. V. & Abel, T. Protein synthesis is required for the enhancement of long-term potentiation and long-term memory by spaced training. J. Neurophysiol. 87, 2770-2777 (2002). 52. Kandel, E. R. The molecular biology of memory storage: a dialogue between genes and synapses. Science 294, 1249-1254 (2001). 53. Silva, A. J., Kogan, J. H., Frankland, P. W. & Kida, S. CREB and memory. Annu. Rev. Neurosci. 21, 127-148 (1998). 54. Liu, R. Y., Fioravante, D., Shah, S. & Byrne, J. H. cAMP response element-binding protein 1 feedback loop is necessary for consolidation of long-term synaptic facilitation in Aplysia. J. Neurosci. 28, 1970-1976 (2008). 55. Liu, R. Y., Cleary, L. J. & Byrne, J. H. The requirement for enhanced CREB1 expression in consolidation of long-term synaptic facilitation and long-term excitability in sensory neurons of Aplysia. J. Neurosci. 31, 6871-6879 (2011a). 56. Liu, R. Y., Shah, S., Cleary, L. J. & Byrne, J. H. Serotonin- and training-induced dynamic regulation of CREB2 in Aplysia. Learn. Mem. 18, 245-249 (2011b). 28 57. Frost, W. N., Castellucci, V., Hawkins, R. D., & Kandel, E. R. Monosynaptic connections made by the sensory neurons of the gill- and siphon-withdrawal reflex in Aplysia participate in the storage of long-term memory for sensitization. Proc. Natl. Acad. Sci. USA 82, 8266- 8269 (1985). 58. Pinsker, H. M., Hening, W. A., Carew, T. J., & Kandel, E. R. Long-term sensitization of a defensive withdrawal reflex in Aplysia. Science 182, 1039-1042 (1973). 59. Beckinschtein, P., Cammarota, M., Igaz, L. M., Bevilaqua, L. R., Izquierdo, I. & Medina, J. H. Persistence of long-term memory storage requires a late protein synthesis- and BDNF- dependent phase in the hippocampus. Neuron 53, 261-277 (2007). 60. Bambah-Mukku, D., Travaglia, A., Chen, D. Y., Pollonini, G. & Alberini, C M. A positive loop via C/EBPβ mediates hippocampal memory autoregulatory BDNF consolidation. J. Neurosci. 34, 12547-12559 (2014). feedback This study characterizes the dynamics of a molecular positive feedback loop demonstrated to be important for the consolidation of long-term (24 h or more) memory, as well as the late repression of Bdnf that terminates the feedback. 61. Genoux, D., Haditsch, U., Knobloch, M., Michalon, A., Storm, D. & Mansuy, I. M. Protein phosphatase 1 is a molecular constraint on learning and memory. Nature 418, 970-974 (2002). 62. Farah, C. A., Weatherill, D., Dunn, T. W., & Sossin, W. S. PKC differentially translocates during spaced and massed training in Aplysia. J. Neurosci. 29, 10281-10286 (2009). 63. Tully, T., Preat, T., Boynton, S. & Del Vecchio, M. (1994) Genetic dissection of consolidated memory in Drosophila melanogaster. Cell 79, 35-47. 64. Yin, J. C., Del Vecchio, M., Zhou, H. & Tully, T. CREB as a memory modulator: induced expression of a dCREB2 activator isoform enhances long-term memory in Drosophila. Cell 81, 107-115 (1995). 65. Smolen, P., Baxter, D. A. & Byrne, J. H. Frequency selectivity, multistability, and oscillations emerge from models of genetic regulatory systems. Am. J. Physiol. 274, C531-542 (1998). 66. Philips, G. T., Tzvetkova, E. I, & Carew, T. J. Transient mitogen-activated protein kinase activation is confined to a narrow temporal window required for the induction of two-trial long term memory in Aplysia. J. Neurosci. 27, 13701-13705 (2007). 29 67. Seese, R. R., Wang, K., Yao, Y. Q., Lynch, G. & Gall, C. M. Spaced training rescues memory and ERK1/2 signaling in fragile X syndrome model mice. Proc. Natl. Acad. Sci. USA 111, 16907-16912 (2014). 68. Rosenblum, K., Futter, M., Voss, K., Erent, M., Skehel, P. A., French, P., Obosi, L., Jones, M. W. & Bliss, T. V. The role of extracellular regulated kinases I/II in late-phase long-term potentiation. J. Neurosci. 22, 5432-5441 (2002). 69. Ajay, S. M., & Bhalla, U. S. A propagating ERKII switch forms zones of elevated dendritic activation correlated with plasticity. HFSP J. 1, 49-66 (2007). 70. Chen, C., Wu, J., Lin, H., Pai, T., Fu, F., Wu, C., Tully, T. & Chiang, A. Visualizing long-term memory formation in two neurons of the Drosophila brain. Science 335, 678-685 (2012). 71. Cepeda, N. J., Coburn, N., Rohrer, D., Wixted, J. T., Moser, M. C. & Pashler, H. Optimizing distributed practice: theoretical analysis and practical implications. Exp. Psychol. 56, 236- 246 (2009). 72. Wei, Y. & Koulakov, A. A. Long-term memory stabilized by noise-induced rehearsal. J. Neurosci. 34, 15804-15815 (2014). 73. Cui, Z., Wang, H., Tan, Y., Zaia, K. A., Zhang, S. & Tsien, J. Z. Inducible and reversible NR1 knockout reveals a crucial role of the NMDA receptor in preserving remote memories in the brain. Neuron 41, 781-793 (2004). This study demonstrated the key point that ongoing neural activity and NMDA receptor activation are essential for long-lasting preservation of stored memory. 74. Tetzlaff, C., Kolodziejski, C., Timme, M., Tsodyks, M. & Worgotter, F. Synaptic scaling enables dynamically distinct short- and long-term memory formation. PLoS Comput. Biol. 9, e1003307 (2013). 75. Miller, P., Zhabotinsky, A. M., Lisman, J. E. & Wang, X. J. The stability of a stochastic CaMKII switch: dependence on the number of enzyme molecules and protein turnover. PLoS Biol. 3, e107 (2005). 76. Pettigrew, D. B., Smolen, P., Baxter, D. A. & Byrne, J. H. Dynamic properties of regulatory motifs associated with induction of three temporal domains of memory in Aplysia. J. Comput. Neurosci. 18, 163-181 (2005). 30 77. Shema, R., Sacktor, T. C. & Dudai, Y. Rapid erasure of long-term memory associations in the cortex by an inhibitor of PKM ζ. Science 317: 951-953 (2007). 78. Volk, L. J., Bachman, J. L., Johnson, R., Yu, Y. & Huganir, R. L. PKM ζ is not required for hippocampal synaptic plasticity, learning and memory. Nature 493, 420-423 (2013). 79. Fioriti, L., Myers, C., Huang, Y. Y., Li, X., Stephan, J. S., Trifillief, P., Colnaghi, L., Kosmidis, S., Drisaldi, B., Pavlopoulos, E., & Kandel, E. R. The persistence of hippocampal-based memory requires protein synthesis mediated by the prion-like protein CPEB3. Neuron 86, 1433-1448 (2015). 80. Ajay, S. M. & Bhalla, U. S. A role for ERKII in synaptic pattern selectivity on the time scale of minutes. Eur. J. Neurosci. 20, 2671-2680 (2004). 81. Frey, U. & Morris R. G. Synaptic tagging: implications for late maintenance of hippocampal long-term potentiation. Trends Neurosci. 21, 181-188 (1998). 82. Babayan, A. H., Krámar, E. A., Barrett, R. M., Jafari, M., Häettig, J., Chen, L. Y., Rex, C. S., Lauterborn, J. C., Wood, M. A., Gall, C. M. & Lynch, G. Integrin dynamics produce a delayed stage of long-term potentiation and memory consolidation. J. Neurosci. 32, 12854- 12861 (2012). 83. Callan, D. E. & Schweighofer, N. Neural correlates of the spacing effect in explicit verbal semantic encoding support the deficient-processing theory. Hum. Brain Mapp. 31, 645-659 (2010). 84. Karpicke, J. D. & Roediger, H. L. Is expanding retrieval a superior method for learning text materials? Mem. Cogn. 38, 116-124 (2010). 85. Zhang, Y., Liu, R. Y., Heberton, G. A., Smolen, P., Baxter, D. A., Cleary, L. J. & Byrne, J. H. Computational design of enhanced learning protocols. Nat. Neurosci. 15, 294-297 (2012). This study found that computational models describing long-term synaptic strengthening can successfully predict optimal spaced training protocols, with irregular inter-trial intervals, that enhance learning. 86. Bhalla, U. S. & Iyengar, R. Emergent properties of networks of biological signaling pathways. Science 283, 381-387 (1999). 87. Kim, M., Huang, T., Abel, T. & Blackwell, K. T. Temporal sensitivity of protein kinase A activation in late-phase long term potentiation. PLoS Comput. Biol. 6, e1000691 (2010). 31 88. Smolen, P., Baxter, D. A. & Byrne, J. H. A model of the roles of essential kinases in the induction and expression of late long-term potentiation. Biophys. J. 90, 2760-2775 (2006). 89. Aslam, N. & Shouval, H. Z. Regulation of cytoplasmic polyadenylation can generate a bistable switch. BMC Syst. Biol. 6, 12-20 (2012). 90. Liu, R. Y., Zhang, Y., Baxter, D. A., Smolen, P., Cleary, L. J. & Byrne, J. H. Deficit in long- term synaptic plasticity is rescued by a computationally predicted stimulus protocol. J. Neurosci. 33, 6944-6949 (2013). 91. Montarolo, P. G., Goelet, P., Castellucci, V. F., Morgan, J., Kandel, E. R. & Schacher, S. A critical period for macromolecular synthesis in long-term heterosynaptic facilitation in Aplysia. Science 234, 1249-1254 (1986). 92. Müller, U. & Carew, T. J. Serotonin induces temporally and mechanistically distinct phases of persistent PKA activity in Aplysia sensory neurons. Neuron 21, 1423-1434 (1998). 93. Korzus, E., Rosenfeld, M. G & Mayford M. CBP histone acetyltransferase activity is a critical component of memory consolidation. Neuron 42, 961-972 (2004). 94. Park, E., Kin, Y., Ryu, H., Kowall, N. W., Lee, J. & Ryu, H. Epigenetic mechanisms of Rubinstein-Taybi syndrome. Neuromol. Med. 16, 16-24 (2014). 95. Alarcon, J. M., Malleret, G., Touzani, K., Vronskaya, S., Ishii, S., Kandel, E. R. & Barco, A. Chromatin acetylation, memory, and LTP are impaired in CBP+/- mice: a model for the cognitive deficit in Rubinstein-Taybi syndrome and its amelioration. Neuron 42, 947-959 (2004). 96. Alberini, C. M. Transcription factors in long-term memory and synaptic plasticity. Physiol. Rev. 89, 121–145 (2009). 97. Zhou, L., Zhang, Y., Liu, R. Y., Smolen, P., Cleary, L. J. & Byrne, J. H. Rescue of impaired long-term facilitation at sensorimotor synapses of Aplysia following siRNA knockdown of CREB1. J. Neurosci. 35, 1617-1626 (2015). 98. Abel, T. & Nguyen, P. V. Regulation of hippocampus-dependent memory by cyclic AMP- dependent protein kinase. Prog. Brain Res. 169, 97-115 (2008). 99. Bartsch, D., Casadio, A., Karl, K. A., Serodio, P. & Kandel, E. R. CREB1 encodes a nuclear activator, a repressor, and a cytoplasmic modulator that form a regulatory unit critical for long-term facilitation. Cell 95, 211-223 (1998). 32 100. Kida, S. A functional role for CREB as a positive regulator of memory formation and LTP. Exp. Neurobiol. 21, 136-140 (2012). 101. Bartsch, D., Ghirardi, M., Skehel, P. A., Karl, K. A., Herder, S. P., Chen, M., Bailey, C. H. & Kandel, E. R. Aplysia CREB2 represses long-term facilitation: relief of repression converts transient facilitation into long-term functional and structural change. Cell 83, 979-992 (1995). 102. Dong, C., Upadhya, S. C., Ding, L., Smith, T. K. & Hegde, A. N. Proteosome inhibition enhances the induction and impairs the maintenance of late-phase long-term potentiation. Learn. Mem. 15, 335-347 (2008). 103. Pasini, S., Corona, C., Liu, J., Greene, L. A., & Shelanski, M. L. Specific downregulation of hippocampal ATF4 reveals a necessary role in synaptic plasticity and memory. Cell Reports 11, 183-191 (2015). 104. Smolen, P., Baxter, D. A. & Byrne, J. H. Simulations suggest pharmacological methods for rescuing long-term potentiation. J. Theoret. Biol. 360, 243-250 (2014). 105. Cuadrado-Tejedor, M., Garcia-Barroso, C., Sanzhez-Arias, J., Mederos, S., Rabal, O., Ugarte, A., Franco, R., Pascual-Lucas, M., Segura, V., Perea, G., Oyarzabal, J., & Garcia- Osta, A. Concomitant histone deacetylase and phosphodiesterase 5 inhibition synergistically prevents the disruption in synaptic plasticity and it reverses cognitive impairment in a mouse model of Alzheimer's disease. Clin. Epigenetics 7, 108 (2015). 106. Ikegaya, Y., Aaron, G., Cossart, R., Aronov, D., Lampl, I., Ferster, D. & Yuste, R. Synfire chains and cortical songs: temporal modules of cortical activity. Science 304, 559-564 (2004). 107. Miller, J. E., Ayzenshtat, I., Carrillo-Reid, L. & Yuste, R. Visual stimuli recruit intrinsically generated cortical ensembles. Proc. Natl. Acad. Sci. USA 111, E4053-4061 (2014). 108. Karlsson, M. P. & Frank, L. M. Awake replay of remote experiences in the hippocampus. Nat. Neurosci. 12, 913-919 (2009). 109. Wu, X. & Foster, D. J. Hippocampal replay captures the unique topological structure of a novel environment. J. Neurosci. 34, 6459-6469 (2014). 110. Hsiang, H. L., Epp, J. R., van den Oever, M. C., Yan, C., Rashid, A. J., Insel, N., Ye, L., Niibori, Y., Deisseroth, K., Frankland, P. W. & Josselyn, S. A. Manipulating a "cocaine engram" in mice. J. Neurosci. 34, 14115-14127 (2014). 33 111. Lynch, G., Cox, C. D. & Gall, C. M. Pharmacological enhancement of memory or cognition in normal subjects. Front. Syst. Neurosci. 8, doi: 10.3389/fnsys.2014.00090 (2014). Acknowledgements This work was supported by National Institutes of Health grants NS073974 and NS019895. 34
1210.6317
1
1210
2012-10-23T18:21:16
On the geometric structure of fMRI searchlight-based information maps
[ "q-bio.NC", "q-bio.QM", "stat.AP", "stat.ML" ]
Information mapping is a popular application of Multivoxel Pattern Analysis (MVPA) to fMRI. Information maps are constructed using the so called searchlight method, where the spherical multivoxel neighborhood of every voxel (i.e., a searchlight) in the brain is evaluated for the presence of task-relevant response patterns. Despite their widespread use, information maps present several challenges for interpretation. One such challenge has to do with inferring the size and shape of a multivoxel pattern from its signature on the information map. To address this issue, we formally examined the geometric basis of this mapping relationship. Based on geometric considerations, we show how and why small patterns (i.e., having smaller spatial extents) can produce a larger signature on the information map as compared to large patterns, independent of the size of the searchlight radius. Furthermore, we show that the number of informative searchlights over the brain increase as a function of searchlight radius, even in the complete absence of any multivariate response patterns. These properties are unrelated to the statistical capabilities of the pattern-analysis algorithms used but are obligatory geometric properties arising from using the searchlight procedure.
q-bio.NC
q-bio
On the geometric structure of fMRI searchlight-based information maps Shivakumar Viswanathan1, Matthew Cieslak, Scott T. Grafton Department of Psychological and Brain Sciences, University of California, Santa Barbara, California, USA Abstract 2 1 0 2 t c O 3 2 ] . C N o i b - q [ 1 v 7 1 3 6 . 0 1 2 1 : v i X r a Information mapping is a popular application of Multivoxel Pattern Analysis (MVPA) to fMRI. Information maps are constructed using the so called searchlight method, where the spherical multivoxel neighborhood of every voxel (i.e., a searchlight) in the brain is evaluated for the presence of task-relevant response patterns. Despite their widespread use, information maps present several challenges for interpretation. One such chal- lenge has to do with inferring the size and shape of a multivoxel pattern from its signature on the information map. To address this issue, we formally examined the geometric basis of this mapping relationship. Based on geometric considerations, we show how and why small patterns (i.e., having smaller spatial extents) can produce a larger signature on the information map as compared to large patterns, independent of the size of the searchlight radius. Furthermore, we show that the number of informative searchlights over the brain increase as a function of searchlight radius, even in the complete absence of any multivariate response pat- terns. These properties are unrelated to the statistical capabilities of the pattern-analysis algorithms used but are obligatory geometric properties arising from using the searchlight procedure. Keywords: MVPA, searchlight, FMRI, pattern-classification 1. Introduction In FMRI, a functional map is an important represen- tation of how cognitive function is related to neu- roanatomy. Such maps provide a topographic rep- resentation of the brain regions that are (and are not) systematically responsive to differing values of a cognitive variable. The size, shape, number and location of the “blobs” (i.e., voxel-clusters meeting some statistical relevance criterion) on the functional maps are the basis for inferences about the neural substrates of the cognitive process. Given the impor- tance of functional maps, there is a continuing need to scrutinize the sensitivity, precision and technical assumptions of the mapping procedure itself. The topic of the current technical note is the mapping procedure used to generate the widely used informa- tion maps (Kriegeskorte et al., 2006). The motivation for information mapping is the sta- tistical concern that a region’s responses to the cog- nitive variable under study might take a complex 1Corresponding author: [email protected] Preprint submitted to arXiV multivariate form. For example, a group of mul- tiple voxels might conjointly respond in a task- relevant manner even though individual voxels may not detectably do so (Haxby et al., 2001; Cox and Savoy, 2003; Haynes, 2006; Norman et al., 2006; Mur et al., 2008; Tong, 2010; Wagner and Rissman, 2010; Formisano and Kriegeskorte, 2012; Serences and Saproo, 2012). Such distributed response pat- terns might be effectively undetectable with conven- tional univariate statistical tests restricted to indi- vidual voxel responses, but detectable with an ex- plicit multivariate test for multivoxel response pat- terns, i.e., using some Multivoxel Pattern Analysis (MVPA) technique. To address this concern in the context of functional mapping, Kriegeskorte et al. (2006) proposed a simple procedure to enable so- phisticated MVPA methods to be readily applied to detect and map brain regions that contain informa- tion about the experimental conditions, irrespective of whether the informative responses are univariate or multivariate. In the proposed procedure, the unit of evaluation is not the single voxel but a “searchlight” – the group of May 3, 2014 voxels contained in a spherical neighborhood of ra- dius r around a single voxel. The searchlight statis- tic is a measure of whether the conjoint responses of this group of voxels contain information about the experimental conditions being tested. Based on these abstractions, an information map is gen- erated as follows: the searchlight statistic is evalu- ated for searchlights centered at every voxel in the brain; and the statistic’s value for each searchlight is mapped to the central voxel of that searchlight. The resulting topographic representation generated by the searchlight-procedure has been referred to as the information map. Such searchlight-based infor- mation maps are now routinely reported in studies that employ MVPA methods (for example, Haynes et al., 2007; Soon et al., 2008; Johnson et al., 2009; Poldrack et al., 2009; Chadwick et al., 2010; Ooster- hof et al., 2010; Nestor et al., 2011; Alink et al., 2011; Golomb and Kanwisher, 2011; Peelen and Kastner, 2011; Stokes et al., 2011; Woolgar et al., 2011; Mor- gan et al., 2011; Oosterhof et al., 2012; Connolly et al., 2012; Kaplan and Meyer, 2012). Notwithstanding their popularity, interpreting a searchlight-based information map presents a va- riety of challenges (Kriegeskorte et al., 2006; Pol- drack et al., 2009; Pereira and Botvinick, 2011; Jimura and Poldrack, 2012). One such challenge is posed by the topographic ambiguity of the information map. Recall that a searchlight statistic computed on the responses of an entire multivoxel searchlight is mapped to a single voxel on the information map, namely, that searchlight’s central voxel. This map- ping protocol is applied to searchlights across the brain irrespective of the number or the spatial loca- tions of the information-carrying voxels within each searchlight. Consequently, the spatial position of an informative voxel on the information maps is a coarse index to the actual location of the informative “pattern” within that voxel’s searchlight. Further- more, since a searchlight has a unique central voxel v, an informative voxel on the information map is not indicative of the actual number of voxels con- stituting the informative pattern within that voxel’s searchlight neighborhood. Given these properties of the information map, we asked: what, if anything, can be reliably inferred about the size and shape of a multivoxel pattern from its corresponding signature on the information map? Previous studies have treated this question as a qualitative concern requiring cautious interpreta- tion. Nonetheless, here we show that information maps are in fact subject to several crisply quantita- tive geometric constraints that strongly govern how such maps can be interpreted. Our analytical results are based on a simple geomet- ric intuition. Since a multivoxel searchlight is de- fined at every voxel across the brain, searchlights centered at different voxels systematically overlap each other, i.e., have voxels in common. Using over- lapping searchlights is crucial to obtain a continuous topographic coverage especially when the locations and spatial extents of voxel-neighborhoods that are task-responsive are unknown a priori. We observed that due to these overlaps, multiple searchlights would be deemed informative merely by virtue of sharing the same task-relevant multivoxel response patterns. Thus we reasoned that the size and shape of a multivoxel group G’s signature on the informa- tion map should be defined by exactly those vox- els which have searchlight-neighborhoods that con- tain G. Using this observation and simple geometric reasoning, we formally deduce some key properties of the relationship between an informative pattern and its corresponding signature on the information map. Based on our formal analysis, we prove here that, for any searchlight radius, a single task-responsive voxel produces a larger signature on the informa- tion map as compared to a distributed multivoxel response pattern. Furthermore, the number of in- formative searchlights over the brain can increase as a function of searchlight radius, without necessar- ily revealing any new information and even in the complete absence of any multivariate response pat- terns. Importantly, these properties are largely inde- pendent of the type of machine-learning algorithm or the testing protocol used to compute the search- light statistic. 2. Model 2.1. Definition: The searchlight decomposition The basis of the searchlight analysis is the geomet- ric structure of the voxel-space in which the brain images are defined. The voxel-space V is defined here as the set of all voxels V augmented with a ge- ometric structure defining the relative spatial posi- tion of the voxels in V , and a distance measure be- tween these voxels. For analytical convenience, we treat the voxel-space as being uniform and connected as described below. A d-dimensional voxel-space V is deemed to be uni- form if every voxel has a neighboring voxel in all d principal directions. Additionally, we assume that the voxel-space V is connected. Specifically, there is a path connecting every pair of voxels vi and vj in V with a path defined here to be an ordered se- quence of voxels (cid:104)vi,··· , vk, vk+1, ....vj(cid:105) where voxel vk+1 is a neighbor of vk along one of the d princi- pal directions. These simplifying assumptions are intended to emphasize the general geometric princi- ples entailed by the searchlight method while delib- erately ignoring the special cases associated with (i) the boundaries of V where a searchlight may be trun- cated; and (ii) distinctions between gray-matter and white-matter voxels and any masking of the latter from the searchlights. Although we refer to search- lights as being volumes in a voxel-space having di- mensionality d = 3, the properties derived here are agnostic to the specific value of d and apply to sur- faces (d = 2) where the searchlights are discs (as in, Oosterhof et al., 2011; Chen et al., 2011). The key abstraction defined by the searchlight method is a decomposition of V into subsets of voxels based on a geometric criterion. Given a voxel space V, we define a searchlight voxel- decomposition using the following indexing func- tion S : V × R → P(V ) (1) where P(V) is the powerset of V, namely, the set of all subsets of V. This indexing function S takes two inputs – the identity of a voxel v in the voxel-space V, and a real-value r ∈ R specifying the search- light’s radius. The searchlight indexing function uses these parameters in conjunction with the geo- metric structure of V to extract and output a set of voxels S(r, v) ∈ P(V ). A voxel v(cid:48) ∈ V is a mem- ber of S(r, v) if and only if the distance between v(cid:48) and v is less than or equal to r. For convenience, we henceforth write Sr(v) to denote the searchlight S(r, v). The resulting searchlight voxel-decomposition of V for a given radius r is defined as Sr(V) = {Sr(v) for all v ∈ V} (2) For clarity, we restrict our usage of the term “search- light” to the cases when the value of the radius of a searchlight r is such that each Sr(V ) is a multi- voxel entity that is not identical with V, that is, 1 < Sr(v) (cid:28) V, for any Sr(v) ∈ Sr(V). We refer to the univariate case where Sr(v) = 1 as the univoxel decomposition. A schematic of the searchlight indexing scheme is shown in Figure 1. 2.2. Definition: Informativeness function The searchlight statistic is a measure of whether the voxels in the searchlight, as a unit, exhibit differ- ences in their conjoint responses to the experimen- tal conditions. More generally, it is a measure of whether the searchlight contains information about the experimental condition, i.e., whether the search- light is informative. As with the radius of the search- light, the specific statistical procedure used to com- pute the searchlight statistic is a discretionary choice made by the researcher (for example, see Pereira and Botvinick, 2011). To describe the searchlight statistic in a procedure- independent manner, we use a binary indicator function, which we refer to as the informativeness function, I : P(V ) → {0, 1}. Given a subset of voxels G ∈ P(V ), the function I returns a value of 1 if the responses of G are deemed to be informative; or 0 if they are not, based on some appropriately specified statistical criterion. Evaluating the informativeness function on the re- sponses corresponding to each searchlight Sr(v) in Sr(V) defines the overall information set for a par- ticular radius I(Sr(V)) = {I(Sr(v)) for all Sr(v) in Sr(V)} (3) The information map is the object obtained when the information set defined above is augmented with the geometric structure of the voxel-space V by map- ping the informativeness value of each searchlight I(Sr(v)) to its corresponding central voxel v. The performance measure of interest here is the total number of informative searchlights for a particular Figure 1: Cartoon of searchlight indexing scheme: Left panel shows two example voxels va and vb that are mapped to searchlights Sr(va) (orange) and Sr(vb) (light gray) respectively, having some radius r, as shown in the right panel. Two searchlights can overlap to varying degrees depending on the distance between their center voxels and the radius. Here the overlap is indicated in dark- gray. The searchlight statistic computed on each searchlight is mapped back to the corresponding central voxels to generate the information map (see text). searchlight decomposition V(cid:88) Fr = I(Sr(vi)) (4) i=1 2.3. Linking assumptions Two simple properties link the structure of the searchlight decomposition Sr(V) to the structure of the information set I(Sr(V)). The first property is that, by virtue of the regularity of their shape and relative positioning, searchlights in Sr(V) can overlap. Consequently, the same voxels in V can be included or sampled by multiple search- lights. The second property is that since a sphere is an arbitrarily chosen and regular shape, it is un- likely that every voxel is necessarily task-relevant in every informative searchlight. Consequently, the in- formativeness of the responses in some particular searchlight volume Sr(v) can be alternatively and accurately interpreted as indicating that some group of voxels in that searchlight volume exhibits task- dependent responses. These two properties can be combined as follows. Let G be a group of task- relevant voxels in a searchlight Sr(v), where G ⊆ Sr(v). Since searchlights share voxels, if some other searchlight Sr(v(cid:48)) also contains G, that is G ⊂ Sr(v(cid:48)), then it implies that Sr(v(cid:48)) should also contain task- relevant information as it includes the task-relevant voxels G. Based on this observation, we make two linking assumptions about the behavior of the procedures used to compute the searchlight statistic and hence the informativeness function I. The first is that we restrict the focus of our analysis to the common mul- tivariate procedure that does not include geomet- ric information about the relative spatial positions of the voxels in a searchlight while computing that searchlight’s informativeness. The second is the Su- perset informativeness (SIN) assumption which postu- lates that: Superset informativeness assumption: If a group of voxels G is informative then every searchlight that con- tains G is also informative. That is, according to the SIN assumption, if I(G) = 1 then I(Sr(v)) = 1 for all v ∈ V where G > 0 and G ⊆ Sr(v). Unless otherwise stated, we will over- load the symbol I to denote an informativeness func- tion that explicitly satisfies these two model require- ments. Although the SIN assumption is based is on a sound deduction, the empirical requirement that it poses may not necessarily be satisfied in practice. Specif- ically, even if it is known that I(G) = 1, the sta- tistical procedure used to evaluate informativeness might fail to detect that a searchlight Sr(v) is in- formative even if G ⊂ Sr(v). Such a Type II er- ror (i.e., failing to reject a false null hypothesis that H0 : I(Sr(v)) = 0) might occur for any of a va- !!!"!"#$%&’()*$+$),*-%./-0’"##!!$"##!$$ riety of reasons, for example, the use of an inap- propriate machine-learning algorithm (Pereira and Botvinick, 2011), insufficient power due to a limited number of samples, and so on. In this regard, the SIN assumption treats the multivoxel pattern anal- ysis techniques as being more sensitive and reliable than might actually be the case in practice. That is, the SIN assumption allows us to establish the infor- mation map’s properties in the best-case indepen- dent of the performance idiosyncrasies of the spe- cific multivariate method being used. Proof. Consider a searchlight Sr(va) centered at voxel va. Since a searchlight is defined at every voxel in V (Equation 2), it follows that there is a search- light defined at every voxel in Sr(va). By definition, a voxel vb ∈ V is a member of Sr(va) if and only if the distance between va and vb is less than or equal to the radius r. Since there is a searchlight Sr(vb) centered at vb ∈ Sr(va), and the distance between va and vb is less than or equal to r, it follows that va is a member of searchlight Sr(vb). Therefore, if vb is a member of Sr(va) then va is a member of Sr(vb). 3. Analytical results Our focus of the current section is to establish how the structure of the sampling bias arises from the searchlight decomposition. We first prove that due to the geometric regularities of a searchlight decom- position, single-voxels and multivoxel groups are sampled with different frequencies, i.e., included in a different number of searchlights. Specifically, sin- gle voxels are included in more searchlights than multivoxel groups. This sampling difference is in- dependent of the searchlight radius. We then ex- tend these results to prove that the frequency with which voxel-groups are sampled increases with the radius of the searchlights, irrespective of the number of voxels in the group. Finally, we prove that the in- formation map mirrors these sampling biases in an optimistic manner, i.e., in a manner that is not neces- sarily warranted by the data. 3.1. Single-voxels and multivoxel-groups are sampled with different frequencies The regularity in the shape of the searchlights and their relative positions the voxel-space define a sys- tematic relationship between each voxel v ∈ V and the searchlights in Sr(V) that contain that voxel v. Firstly, if a voxel va is a member of the searchlight Sr(vb), then by symmetry, the voxel vb is a mem- ber of the searchlight Sr(va) (Lemma 1). Secondly, two distinct voxels va and vb are not simultaneously included in every searchlight that contains either of these voxels (Lemma 2). Lemma 1. If a voxel vb is a member of Sr(va) then the voxel va is a member of Sr(vb), where va, vb ∈ V and Sr(va), Sr(vb) ∈ Sr(V. Lemma 2. For any two non-identical voxels va and vb, where va, vb ∈ V and Sr(va) (cid:54)= Sr(vb), there necessar- ily exists a searchlight that contains va but not vb and a different searchlight that contains vb but not va. Proof. This claim can be proved in two steps based on the distance between va and vb. First consider the case where the distance between va and vb is greater than 2r, that is, the diameter of a searchlight. By definition, a searchlight con- tains voxels that have a distance less than or equal to r from that searchlight’s central voxel. Due to the spherical shape of the searchlight, the maximum distance between any two voxels in a searchlight is equal to 2r. If the distance between va and vb is greater than 2r, there does not exist any searchlight of radius r that contains both va and vb as members. Thus, it follows that there exists some searchlight that contains va but not vb; and some other search- light that contains vb but not va. Now consider the second case where the distance be- tween va and vb is less than or equal to 2r. Since the distance between these two voxels is less than the maximum distance between some two voxels in a searchlight, in a uniform voxel-space there necessar- ily exists some searchlight Sr(v) that contains both va and vb as members. Contrary to the proposition, let us assume that both va and vb are contained in ev- ery searchlight that contains either va or vb. That is, if a searchlight Sr(v) contains va, then it necessarily contains vb, and vice versa. Recall that, from Lemma 1, a voxel va is contained in every searchlight Sr(v) where v ∈ Sr(va). Now, based on the contradictory assumption, it implies that vb is also contained in every such searchlight Sr(v) where v ∈ Sr(va). By the same reasoning, va should be contained in every searchlight Sr(v) where v ∈ Sr(vb). If these condi- tions hold true, then it implies that every voxel in Sr(va) is also contained in Sr(vb); and every voxel in Sr(vb) is also contained in Sr(va). If this the case, then the searchlights S(va) and S(vb) are identical as they contain exactly the same voxels. This rela- tionship, however, contradicts the requirement that Sr(va) (cid:54)= Sr(vb). Thus, the assumption that va and vb are both contained in every searchlight that contains either va or vb cannot be true. Therefore, there necessarily exists a searchlight that contains va that does not contain vb, and some other searchlight that contains vb but not va. Armed with the properties described by Lemmas 1 and 2, we can now numerically estimate the num- ber of searchlights that include a given individual voxel. Theorem 3. A voxel v is contained in exactly Nr(v) dif- ferent searchlights, where Nr(v) is the number of voxels contained in the searchlight Sr(v). Proof. From Lemma 1, a voxel v is contained in each searchlight Sr(v(cid:48)), if and only if v(cid:48) is a voxel in Sr(v). Let Nr(v) be the number of voxels in Sr(v). Therefore, v is present in each of these Nr(v) search- lights. For simplicity, we treat Nr(v) as being the same for every searchlight, and write Nr to indicate the canonical number of voxels contained in a spheri- cal volume of radius r, for a given resolution of the voxel-space. From Theorem 3, we see that the radius, a parameter chosen by the researcher, directly specifies how often information in a particular voxel is sampled by mul- tiple searchlights. For voxels of size 3mm × 3mm × 3mm, the number of voxels contained in searchlights of different radii are shown in Figure 2. As can be seen, the number of voxels in a searchlight, that is Nr, grows rapidly with the radius of the searchlight r, and consequently so do the number of searchlights that include a particular voxel. Figure 2: Number of voxels in a searchlight volume (Nr) as a function of searchlight radius (r) in a 3×3×3 mm3 voxel-space. on the relative spatial locations of the voxels in the group. Theorem 4. A group of voxels G containing more than one voxel is contained in strictly less than Nr search- lights. Proof. A voxel is contained in Nr searchlights, from Theorem 3. Consequently, every voxel in G is each contained in Nr searchlights. From Lemma 2, for any two voxels va and vb, there is necessarily a search- light that contains va and not vb, and vice versa. Therefore, of the Nr searchlights containing va, there necessarily exists at least one searchlight that con- tains va but not vb. Thus, the number of searchlights that simultaneously contain both va and vb must be less than Nr. Since G contains multiple voxels, any pair of voxels in G must be simultaneously con- tained in less than Nr searchlights. Therefore, all the voxels in G cannot be simultaneously contained in Nr searchlights, and G must be contained in strictly less than Nr searchlights. 3.2. The sampling frequency of voxel(s) increases with searchlight radius Since searchlights are intended to identify multi- voxel response patterns, we extend the single-voxel property in Theorem 3 to quantify the membership of a group of multiple voxels placing no constraint Although single-voxels and multivoxel groups are included in different numbers of searchlights for any radius r, we now show that the absolute number of searchlights that include either a single-voxel or 456789101112050100150200250Radius (mm)Number of voxels in searchlight ( Nr ) a multivoxel group increases with the radius of the searchlight. Lemma 5. A searchlight of radius ra is fully contained in more than one searchlight of radius rb, where rb > ra and Sra(v) ⊂ Srb(v) for all v ∈ V. Proof. Consider two searchlights centered at the same voxel v – one that has a radius ra, and the other having radius rb. By definition, since Sra(v) ⊂ Srb(v), all the voxels in Sra(v) are members of Srb(v), and there exists at least one voxel in Srb(v) that is not in Sra(v). Now, consider the searchlight Sz(v), having radius z = rb − ra. Due to the spherical shape of search- lights, the maximum distance between a voxel v(cid:48)(cid:48) in Sz(v) and some voxel v(cid:48) in Sra(v) is equal to ra + z = rb. Therefore, all other voxels in Sra(v) must have distances less than or equal to rb. Since the distance between these two maximally dis- tant voxels v(cid:48) and v(cid:48)(cid:48) is equal to rb, the voxel v(cid:48) must be contained in a searchlight of radius rb that is cen- tered at v(cid:48)(cid:48), namely, Srb(v(cid:48)(cid:48)). Since all other vox- els in Sra(v) have a distance less than or equal to rb from v(cid:48)(cid:48), it follows that every voxel in Sra(v) is also contained in the searchlight Srb(v(cid:48)(cid:48)). Thus every voxel in Sra(v) is contained in at least two search- lights having radius rb, namely, Srb(v) and Srb(v(cid:48)(cid:48)). Therefore, a searchlight Sra(v) is contained in more than one searchlight of radius rb, where rb > ra and Sra(v) ⊂ Srb(v). Using Lemma 5, we can now prove a general scaling property. Irrespective of the size of a voxel group, the frequency with which it is sampled by different searchlights increases with the radius of the search- light - a property that we prove next. Theorem 6. A group of voxels G is contained in more searchlights of radius rb than searchlights of radius ra, where G ⊆ Sra, rb > ra and Sra(v) ⊂ Srb(v), for all v ∈ V. Proof. Let Kra and Krb be the number of searchlights of radius ra and rb that contain G. Since Sra(v) ⊂ Srb(v) for every v ∈ V, it follows, by transitivity, that if G ⊆ Sra(vi) for some voxel vi ∈ V, then G ⊆ Srb(vi). Therefore, the number of search- ≮ Kra. lights of radius rb that contain G cannot be strictly less than that for ra, that is, Krb By the transitivity of the subset relation, if G ⊆ Sra(vi) and Sra(vi) ⊂ Srb(vj) for some vi, vj ∈ V, then it follows that G ⊂ Srb(vj). From Lemma 5, a searchlight of radius ra is contained in multiple searchlights of radius rb where rb > ra and Sra(v) ⊂ Srb(v) (for all v ∈ V). Since there is more than one searchlight of radius rb containing Sra(vi), for every searchlight for which G ⊆ Sra(vi) holds true, it im- plies that Krb ≥ Kra. From Theorems 3 and 4, the number of searchlights of radius ra that can contain G is less than or equal to Nra. Consequently, in a uniform and connected voxel-space, it follows that there exist two adjacent voxels vi and vj in V such that G is a subset of Sra(vi) but is not a subset of Sra(vj). From Lemma 5, searchlights of radius rb centered at voxels within rb − ra from vi fully contain all voxels in Sra(vi). Since Sra(v) ⊂ Srb(v), it implies that the distance of vj to vi is less than or equal to rb − ra. Therefore, Sra(vi) ⊂ Srb(vj) and consequently G ⊂ Srb(vj). Since G (cid:54)⊂ Sra(vj) and G ⊂ Srb(vj), it implies that there exists at least one voxel at which a searchlight of radius rb contains G, but where a searchlight of ra- dius ra does not contain G. Consequently, Krb must be strictly greater than Kra, that is, the group of vox- els G is contained in more searchlights of radius rb than ra. Theorem 6 above establishes that the number of searchlights that include either a voxel or group of voxels increases monotonically with the radius of the searchlight. How then does this scaling of the sampling bias influence the properties of the infor- mation map? 3.3. An optimistic bias in the information map Recall that Fr (Equation 4) is an index of the sen- sitivity of the searchlight method in detecting mul- tivoxel response patterns, and is equal to the to- tal number of informative searchlights with a par- ticular search decomposition. We now prove that as a direct consequence of how the sampling bias scales with the searchlight radius, the value of Fr also increases strictly monotonically with increasing searchlight radius. Theorem 7. For two searchlight radii, ra and rb, where rb > ra and Sra(v) ⊂ Srb(v) for every v ∈ V, if 0 < Fra < V then Frb > Fra. Proof. Since Sra(v) ⊂ Srb(v) for every v ∈ V, by the SIN assumption, it follows that if I(Sra(v)) = 1 then I(Srb(v)) = 1, for any voxel v ∈ V. Therefore, the number of informative searchlights of radius rb can- ≮ Fra, not be strictly less than that for ra, that is, Frb for any value of Fra. From Lemma 5, a searchlight of radius ra is con- tained in multiple searchlights of radius rb where rb > ra and Sra(v) ⊂ Srb(v) (for all v ∈ V). For every searchlight for which I(Sra(v)) = 1, there is more than one searchlight of radius rb containing Sra(v). Since each informative searchlight of radius ra is a subset of multiple searchlights of radius rb, by the SIN assumption, it implies that Frb ≥ Fra. Let 0 < Fra < V . Since Fra < V , there neces- sarily exist two adjacent voxels vi and vj such that I(Sra(vi)) = 1 and I(Sra(vj)) = 0. By the same logic used to prove Theorem 6, searchlights of ra- dius rb centered at voxels within rb − ra from vi fully contain all voxels in Sra(vi). Consequently, by the SIN assumption, I(Srb(vj)) = 1. This implies that a searchlight centered at voxel vj is informative if it has a radius rb but not if it has a radius ra. Therefore Frb > Fra. What does Theorem 7 have to do with optimism? The monotonic increases in the number of informa- tive searchlights is due to increases in the sampling bias, which in turn is due to the use of a multivoxel searchlight. Specifically, it is possible to obtain an increased “sensitivity” of the information map sim- ply by increasing the radius of the multivoxel search- lights, with no reference to the statistical properties of the voxel-responses, i.e., whether they in fact ex- hibit multivariate response differences. 4. An illustration In this section, we present simulations to provide a concrete intuition for the analytical results above, and their implications. For ease of demonstration, the voxel-space V for all simulations consisted of a single axial slice having two principal directions. All the voxels in this voxel-space were populated with simulated response information from two fictitious experimental conditions A and B. These simulated data were subjected to the searchlight-procedure to produce information maps. The radius r of the searchlights used for the searchlight decomposition Sr(V) was varied systematically to produce a cor- responding information map for each radius value. The radius took the values: 4 mm, 6 mm, 8 mm, 10 mm and 12 mm, corresponding to searchlights con- taining 5 voxels, 13 voxels, 21 voxels, 37 voxels and 49 voxels respectively. The simulated response-data differed in the num- ber and relative spatial location of the voxels that were responsive to the experimental conditions. In the first of these simulations discussed next, a single voxel contained task-relevant information while all the remaining voxels did not. 4.1. The needle-in-the-haystack effect Suppose there exists some voxel in V, say v0, that ex- hibits a response difference to the experimental con- ditions such that the informativeness function iden- tifies v0 as being task-relevant, that is, I(v0) = 1. Since I(v0) = 1, by the SIN assumption it follows that each of the searchlights that contain v0 should also be deemed to be informative as well. Recall that, according to Theorem 3, each voxel v in V is contained in exactly Nr searchlights where Nr = Sr(v). It then follows that the signal-carrying voxel v0 should be contained in Nr searchlights, each be- ing centered at a voxel in Sr(v0). Thus, a single signal-carrying voxel (a “needle”) should produce a cluster having Nr voxels on the information map (a “haystack”). To simulate this “needle-in-the-haystack” effect, the task-relevant responses of v0 in conditions A and B took the form illustrated in Figure 3(a). The re- sponses to both conditions were drawn randomly from a normal distribution with standard deviation σ = 1. The voxel v0’s mean response to condition A was µA = +2; and µB = −2 for condition B. The responses of all other (non task-relevant) voxels were drawn from normal distributions having σ = 1 where µA = µB = 0. To maximize the sensitivity of the searchlight statistic and emulate the require- ments of the SIN assumption, a total of 300 samples were drawn for each condition. The spatial position of voxel v0 is shown in blue in Figure 3(b). The voxel (a) (b) Figure 3: Single-voxel response: (a) Scatter plot of simulated responses of voxel v0 to conditions A and B. A total of 300 samples were drawn per condition (see text). (b) Spatial location of voxel v0 (indicated by blue square) in the simulated single-slice voxel space. was placed far from the boundaries of the slice to avoid truncations of the searchlights and to emulate a uniform voxel-space in the vicinity of v0. With this setup, the searchlight decomposition and testing procedure was implemented using the PyMVPA toolbox (Hanke et al., 2009). Each search- light’s informativeness was determined by evaluat- ing the decodability of its responses, i.e., testing for the existence of a model that accurately classifies a sample’s membership in each condition based on the searchlight’s responses (Pereira and Botvinick, 2011; Pereira et al., 2009). Decodability was tested using a linear Support Vector Machine (SVM) with a soft-margin regularization parameter, C = 1. The searchlight statistic was the mean classification accu- racy obtained using a Leave-One-Out (LOO) cross- validation procedure. Figure 4(a) shows the information maps obtained (thresholded at 60%). In the upper-panel, going from the left to the right in order of increasing radius, we see that there is a single high accuracy cluster (red- colored voxels) centered at the signal-carrying voxel v0, and this cluster grows in size with increasing radius. The lower-panel shows an expanded view of this high accuracy cluster, thresholded at 80%. Consistent with the predictions described above, for each radius, the size and shape of these clusters on the information map correspond exactly to the size, shape and location of the searchlight Sr(v0) centered at voxel v0. Furthermore, consistent with Theorem 7, the number of informative searchlights identified (Fr) increases in a monotonic manner with the ra- dius of the searchlight, even though there is no dif- ference in the actual information present or even any multivoxel response patterns to speak of. Figure 4(b) shows the values on the information map from a single 1D segment running horizontally through the voxel v0 through the diameter of the searchlights centered at v0. The voxel v0 is assigned a value 0. Consistent with the SIN assumption, the accuracies on the information map do not exhibit a smooth degradation as a function of the distance from v0. Critically, this pedestal-like profile is unlike the profile that would be expected if the searchlights were the equivalent of a “spatial smoothing” kernel on the information map. What is the comparable effect on the information map when the task-relevant signal is distributed over multiple voxels? We next consider this sce- nario. 4.2. The haystack-in-the-needle effect Suppose there are two voxels, v1 and v2 in V, that conjointly exhibit a response difference to the ex- perimental conditions. However, neither voxel by itself shows a task-relevant difference. That is, I({v1, v2}) = 1 and I(v1) = I(v2) = 0. By the SIN assumption, every searchlight that contains both v1 and v2 should be informative, but searchlights that AB−5−4−3−2−1012345ConditionResponse of voxel v0!! (a) (b) Figure 4: Information maps for single-voxel response: (a) Upper panel shows the information-maps across the entire slice as a function of increasing searchlight radius going from left to right (threshold = 60% accuracy). Lower panel shows the corresponding expanded view of the high-accuracy clusters (threshold = 80% accuracy) for each searchlight radius value. Horizontal lines are to provide a common reference to compare relative sizes of clusters. (b) Cross-sectional profile of the information map showing classification accuracies for searchlights centered at voxels on a horizontal 1D slice through the voxel v0. The dotted horizontal line indicates the thresholding value of 80%. Only the profiles for radii 4mm, 8mm and 12mm are shown. contain either v1 or v2 alone would not necessarily be informative. Recall that, according to Theorem 4, a group of multiple voxels (i.e., having more than one voxel) is contained in strictly less than Nr search- lights. It then follows that the signal-carrying voxel group {v1, v2} should produce a cluster having less than Nr voxels on the information map, i.e., a mul- tivoxel “haystack” should produce a “needle”-like cluster, unlike the needle-in-the-haystack scenario in Section 4.1 above. To simulate this “haystack-in-the-needle” effect, the task-relevant responses in the two voxels v1 and v2 took the form shown in Figure 5(a). The responses to each condition were drawn randomly from a normal distribution having standard deviation σ = 1. Each voxel’s mean response to conditions A and B are shown as dotted lines. The voxel v1 had an identical mean response to both conditions A and B, specif- ically, µA = µB = 0 (the horizontal dotted line); while voxel v2’s mean response to condition A was µA = +0.5 and to condition B was µB = −0.5 (in- dicated by each of the dotted vertical lines). Impor- tantly, the responses of voxel v1 and v2 to both con- ditions were correlated negatively. The response of voxel v1 on condition A, denoted as X1,A was equal to −X2,A, the response of voxel v2 to condition A. Similarly, for condition B, X1,B = −X2,B. The simu- lated responses of all other voxels were drawn from distributions having σ = 1 and µA = µB = 0, and were uncorrelated with the responses in either voxel v1 or v2. As with the previous simulation above, a total of 300 samples were drawn for each condition. With signals of this form, the conjoint responses of voxels v1 and v2 to conditions A and B are linearly separable (see Figure 5(a)). However, A and B can- not be distinguished from the responses in v1, but should be weakly discriminable from the responses in v2. The relative spatial positions of v1 and v2, indicated as blue squares, are shown in Figure 5(b). We con- sidered two cases, where v1 and v2 were separated by 2 voxels in one case; and by 3 voxels in the other. When v1 and v2 have a separation of 2 voxels, there is no one searchlight of radius 4mm that can con- tain both of these voxels. With a separation of 3 voxels, there are no searchlights of radius 4 mm, 6 mm, or 8 mm that can contain both v1 and v2. With this setup, the searchlight decomposition and testing procedure was simulated in the same manner as in Section 4.1. Figure 6 shows the portions of the information maps INFORMATION MAPSr = 4mmr = 8mmr = 10mmr = 12mmr = 6mm5 voxels21 voxels37 voxels13 voxels49 voxelsAccuracy > 80%Accuracy > 60%60%100%−10−6 −2 2 6 10 405060708090100Voxel indexClassification accuracy (%) 4mm8mm12mm (a) (b) Figure 5: Multivoxel response: (a) Scatter plot of simulated responses of voxels v1 and v2 to conditions A (gray squares) and B (black dots). Dotted lines indicate the mean response of each voxel alone to A and B. (b) Spatial location of voxels v1 and v2 (indicated by blue squares) in the simulated single-slice voxel space. The separation between v1 and v2 was either 2 or 3 voxels (see text). The responses of voxel v2 show a weak response difference to the conditions, indicated by the dotted circle. in the vicinity of voxels v1 and v2 (thresholded at 60%). In all the information maps, the above- threshold cluster takes the size and shape of the cor- responding searchlight and is centered at voxel v2, namely, the voxel exhibiting a weak response dif- ference to conditions A and B. This “needle-in-the- haystack” organization is consistent with the simu- lations in Section 4.1, and is invariant to the number of voxels separating v1 and v2. Now, observe that the clusters in several, but not all, of the information maps contain sub-clusters con- sisting of voxels having high classification accuracies (indicated in red). These voxels on the information map correspond to the centers of searchlights that contain both v1 and v2. As required by Theorem 4, for each radius, the number of high-accuracy vox- els in the cluster are less than Nr. Due to the geo- metric constraint defined by the separation between v1 and v2, the presence of any high-accuracy voxels at all in an information map depends on the radius of the searchlights used. For example, information maps obtained with searchlights of radius 4 mm do not contain any high-accuracy voxels for both sep- arations (top row), while the information maps for searchlights of radius 8mm contain high-accuracy voxels for the 2 voxel separation but not for the 3 voxel separation. Figures 6(a) and (b) show the 1D cross-section of the information map through the horizontal diame- ter of the clusters for the 2 voxel and 3 voxel separa- tions respectively. As evident, there is a “smearing”, rather than smoothing, of the accuracies with grow- ing radius values, as in Figure 4(b). Furthermore, when a searchlight is large enough to include both v1 and v2, there is a large increase in the classifica- tion accuracy. The above simulations confirm the basic statisti- cal premise motivating the searchlight-procedure, namely, the ability of a multivoxel pattern anal- ysis method to detect distributed response pat- terns. However, for any radius, the size of the clusters produced by multivoxel response patterns are smaller than those produced by single voxel response-differences. Consistent with Theorem 7, the number of informative searchlights identified in- creases in a monotonic manner with the radius of the searchlight. 4.3. Whole-brain inflation maps The previous two simulations demonstrated signal- dependent effects caused by the sampling bias inher- ent in the searchlight decomposition. However, ac- cording to Theorem 7, there should be a monotonic increase in the number of informative searchlights as a function of radius, irrespective of the actual distri- bution of task-relevant voxels/voxel-groups across the brain. This monotonic scaling of the size of the “blobs” on the information map makes plausible a rather unusual scenario – an information map where −3−2−10123−3−2−10123Response of voxel v2Response of voxel v1 Condition ACondition B!!!"!"#$%$&’() (a) (b) (c) Figure 6: Information maps for multivoxel response: (a) Left and right panels show portions of information map (threshold = 60% accuracy) centered around voxels v1 and v2 as a function of increasing searchlight radius (top to bottom); and the separation between v1 and v2 – 2 voxel separation (left panel), and 3 voxel separation (right panel). Plots (b) and (c): Cross-sectional profile of the information map showing classification accuracies for searchlights centered at voxels on a horizontal 1D slice through the voxels v1 and v2 for separation of 2 voxels (b), and 3 voxels (c). The lower dotted horizontal line indicates the thresholding value of 60%, and the upper dotted line indicates the regime of the high-accuracy searchlights (80%). !"#$""#!!"!#$%%!!"!#&%%!!"!’%%!!"!(%%!!"!)%%&!*+,-./%&’()(*+,-0!*+,-./%&’()(*+,-12345678142!679:−8 −4 0 4 8 12 405060708090100Voxel indexClassification accuracy (%) 4mm8mm12mm−8 −4 0 4 8 12 405060708090100Voxel indexClassification accuracy (%) 4mm8mm12mm every searchlight in the brain is deemed to be infor- mative. This scenario was motivated by results recently re- ported by Poldrack et al. (2009). In that study, in- formation maps were generated using searchlights of radius 4 mm and 8 mm. Rather remarkably, with a radius of 8 mm, only one region in the informa- tion map (the bilateral dorsolateral prefrontal cor- tex) was found to be uninformative while every other searchlight was informative. This whole-brain cov- erage was, however, not the case with the 4 mm searchlights. Given the inflationary relationship be- tween Fr (the number of informative searchlights) and searchlight radius r that established in the pre- vious sections, curiosity asked: could an informative whole-brain arise (i.e., Fr = V) by random chance with a suitably chosen searchlight radius? This question can be formulated as a covering prob- lem. Consider a finite 3D voxel space correspond- ing to one containing the brain, approximated as a cubic volume of size NX × NY × NZ, where Ni is the number of voxels along the principal direction i. Suppose there is a minimum covering set of search- lights Cr ⊂ Sr(V) such that every voxel in V is con- tained in some searchlight in Cr. Recall that a single- voxel signal can produce a cluster having Nr voxels on the information map, due to Theorem 3. If the central voxel of each of the searchlights in Cr was in- formative, it would follow that searchlights centered at every voxel in every one of the searchlights in Cr would also be informative. Since every voxel in V is present in some searchlight in Cr, it implies that a rather sparse distribution of informative single- voxels specified by Cr could produce an information map where every searchlight in Sr(V) would be in- formative (with the proviso that the SIN assumption holds true.) The sparsity of these informative single-voxels can be readily approximated if we use cubical volumes as a proxy for the spherical shape of the searchlight volumes. A cube of side w voxels would fully con- √ tain a sphere having radius r = w/2, and would be fully contained in a sphere of radius w 3/2. With this simplification, the minimal number of search- light cubes required to cover the voxel space V is readily approximated as the volume of the voxel space divided by the volume of each searchlight cube, that is, ≈ (cid:100)(NX NY NZ)/(w3)(cid:101). Figure 7: Number of non-intersecting cubes required to fill a volume of dimensions Nx = 52, Ny = 63, Nz = 45 (Ni = num- ber of voxels in principal direction i) as a function of cube-size. Values next to each data-point indicate the actual number of fill- ing cubes for each cube-size. For voxels of size 3 × 3 × 3 mm3, we approximate the size of the voxel space with the following val- ues NX = 52 voxels, NY = 63 voxels and NZ = 45 voxels. Figure 7 shows the minimum number of cu- bical volumes required to cover V as a function of w, where w took values 1, 3, 5, 7, 9, 11. A searchlight cube of side w = 1 is equivalent to a single voxel so the size of the covering set Cr is equal to the total number of voxels in V, namely, 147420. However, increasing values of w produce a rapid decrease in the size of the covering set. For w = 3 voxels, a cubical volume that would fully contain a spherical searchlight of radius 4 mm, a total of 5460 equally spaced signal-carrying voxels can produce an information map where every searchlight is in- formative. However, for a cubical volume with side w = 7 voxels, corresponding to spherical volumes of radius 8mm, a mere 430 voxels are required for such a fully informative map. Stated differently, an information map with a single task-relevant cluster made up of every voxel in V can be generated from a mere 430 regularly spaced voxels of the 147, 420 vox- els in V, that is, 430 voxels that enable the conditions to be distinguished whether due to the presence of true signal or by random chance. This potential for a small number of single voxels (i.e., ≈ 0.003% 1357911102103104105Cube width w (voxels)Number of cubes covering V (log10)147,4204301112031,1805,460 of V) to drive the structure of the entire informa- tion map simply by the choice of the searchlight ra- dius presents an important consideration for draw- ing neurobiological interpretation. it is maximal when explicitly assuming a highly sen- sitive and robust MVPA technique, namely one sat- isfying the superset informativeness (SIN) assump- tion. 5. Discussion 6. Acknowledgments Knowledge of the actual information-carrying vox- els in each informative searchlight would make the information map irrelevant. These actual informa- tive voxels could be directly reported, hence resolv- ing the over-counting that arises from their inclusion in multiple searchlights. One possible implementa- tion would be to identify task-relevant voxels in each searchlight, and then combine these identified vox- els across searchlights. However, requiring the iden- tification of the actual informative voxels in each searchlight could reduce the generality of the search- light method. When pattern classifiers are used to compute the searchlight statistic, each voxel (or fea- ture) in a searchlight is typically assigned a weight, and the weighted combination of the multivoxel re- sponses is used to make a classification decision. However, the specific basis for assigning weights to individual features is highly dependent on the spe- cific machine learning algorithm and its inductive assumptions (Mitchell, 1980; Wolpert, 1996; Guyon et al., 2002; Pereira et al., 2009). Consequently, ap- propriate techniques would be required to allow re- sults to be compared across studies that use different MVPA-techniques. Until such advances are made, the analytical frame- work described above provides several constraints on alternate interpretations of the information map. Our results present a strong argument against mea- suring the sensitivity of information mapping by a count of the number of informative searchlights. The seemingly high sensitivity of the searchlight method as judged by such a performance measure in part has a rather trivial explanation. Specifically, an expla- nation in the obligatory geometric properties of the searchlight-method as discussed above rather than an explanation related to underlying neural orga- nization, or the sophisticated machine-learning al- gorithms used to analyze multivoxel response pat- terns, or the widely discussed merits of multivariate statistical evaluations. Indeed, the upshot of the op- timistic scaling of this performance measure is that This work was supported by the U.S. Army Re- search Office through the Institute for Collaborative Biotechnologies under Contract No. W911NF-09-D- 0001. 7. References References Alink, A., Euler, F., Kriegeskorte, N., Singer, W., Kohler, A., 2011. Auditory motion direction encoding in auditory cortex and high-level visual cortex. Human Brain Mapping. Chadwick, M. J., Hassabis, D., Weiskopf, N., Maguire, E. A., 2010. Decoding individual episodic memory traces in the hu- man hippocampus. Current Biology 20 (6), 544–7. Chen, Y., Namburi, P., Elliott, L. T., Heinzle, J., Soon, C. S., Chee, M. W. L., Haynes, J.-D., 2011. Cortical surface-based search- light decoding. Neuroimage 56 (2), 582–92. J. S., Gors, J., Hanke, M., Halchenko, Y. O., Wu, Y.-C., Abdi, H., Haxby, J. V., 2012. The representation of biological classes in the human brain. Jour- nal of Neuroscience 32 (8), 2608–18. Connolly, A. C., Guntupalli, Cox, D. D., Savoy, R. L., Jun 2003. Functional magnetic reso- nance imaging (fmri) ”brain reading”: detecting and classi- fying distributed patterns of fmri activity in human visual cortex. Neuroimage 19 (2 Pt 1), 261–70. Formisano, E., Kriegeskorte, N., 2012. Seeing patterns through the hemodynamic veil - the future of pattern-information fMRI. Neuroimage 62 (2), 1249–56. Golomb, J. D., Kanwisher, N., 2011. Higher level visual cortex represents retinotopic, not spatiotopic, object location. Cere- bral Cortex Epub. Guyon, I., Weston, J., Barnhill, S., Vapnik, V., 2002. Gene selec- tion for cancer classification using support vector machines. Machine learning 46 (1), 389–422. Hanke, M., Halchenko, Y. O., Sederberg, P. B., Hanson, S. J., Haxby, J. V., Pollmann, S., 2009. PyMVPA: A python toolbox for multivariate pattern analysis of fMRI data. Neuroinfor- matics 7 (1), 37–53. Haxby, J. V., Gobbini, M. I., Furey, M. L., Ishai, A., Schouten, J. L., Pietrini, P., 2001. Distributed and overlapping represen- tations of faces and objects in ventral temporal cortex. Science 293 (5539), 2425–30. Haynes, J., 2006. Decoding mental states from brain activity in humans. Nature Reviews Neuroscience 7 (7), 523–534. Haynes, J.-D., Sakai, K., Rees, G., Gilbert, S., Frith, C., Passing- ham, R. E., Feb. 2007. Reading Hidden Intentions in the Hu- man Brain. Current Biology 17 (4), 323–328. tions in human visual cortex. Neuroimage 56 (3), 1540–5. Tong, F., Dec. 2010. Pattern Classification Analysis. Annual Re- view of Psychology 63 (1), 110301102248092. Wagner, A. D., Rissman, J., Dec. 2010. Distributed representa- tions in memory: insights from functional brain imaging. An- nual Review of Psychology 63 (1), 110301102248092. Wolpert, D., 1996. The lack of a priori distinctions between learning algorithms. Neural computation 8 (7), 1341–1390. Woolgar, A., Thompson, R., Bor, D., Duncan, J., May 2011. Multi- voxel coding of stimuli, rules, and responses in human fron- toparietal cortex. Neuroimage 56 (2), 744–52. Jimura, K., Poldrack, R. A., 2012. Analyses of regional-average activation and multivoxel pattern information tell comple- mentary stories. Neuropsychologia 50 (4), 544–52. Johnson, J. D., McDuff, S. G. R., Rugg, M. D., Norman, K. A., 2009. Recollection, familiarity, and cortical reinstatement: a multivoxel pattern analysis. Neuron 63 (5), 697–708. Kaplan, J. T., Meyer, K., 2012. Multivariate pattern analysis reveals common neural patterns across individuals during touch observation. Neuroimage 60 (1), 204–12. Kriegeskorte, N., Goebel, R., Bandettini, P., 2006. Information- based functional brain mapping. Proceedings of the National Academy of Sciences 103 (10), 3863–8. Mitchell, T., 1980. The need for biases in learning generaliza- tions. Tech. Rep. CBM-TR-5-110, Department of Computer Science, Rutgers University. Morgan, L. K., Macevoy, S. P., Aguirre, G. K., Epstein, R. A., 2011. Distances between real-world locations are represented in the human hippocampus. Journal of Neuroscience 31 (4), 1238–45. Mur, M., Bandettini, P. A., Kriegeskorte, N., 2008. Revealing representational content with pattern-information fMRI–an introductory guide. Social Cognitive and Affective Neuro- science 4 (1), 101–109. Nestor, A., Plaut, D. C., Behrmann, M., 2011. Unraveling the distributed neural code of facial identity through spatiotem- poral pattern analysis. Proceedings of the National Academy of Sciences 108 (24), 9998–10003. Norman, K. A., Polyn, S. M., Detre, G. J., Haxby, J. V., Sep. 2006. Beyond mind-reading: multi-voxel pattern analysis of fMRI data. Trends in Cognitive Sciences 10 (9), 424–430. Oosterhof, N. N., Tipper, S. P., Downing, P. E., Jan 2012. View- point (in)dependence of action representations: An MVPA study. J Cogn Neurosci. Oosterhof, N. N., Wiestler, T., Downing, P. E., Diedrichsen, J., May 2011. A comparison of volume-based and surface-based multi-voxel pattern analysis. Neuroimage 56 (2), 593–600. Oosterhof, N. N., Wiggett, A. J., Diedrichsen, J., Tipper, S. P., Downing, P. E., Aug 2010. Surface-based information map- ping reveals crossmodal vision-action representations in hu- man parietal and occipitotemporal cortex. Journal of Neuro- physiology 104 (2), 1077–89. Peelen, M. V., Kastner, S., Jul 2011. A neural basis for real-world visual search in human occipitotemporal cortex. Proceedings of the National Academy of Sciences 108 (29), 12125–30. Pereira, F., Botvinick, M., May 2011. Information mapping with pattern classifiers: A comparative study. NeuroImage 56 (2), 476–496. Pereira, F., Mitchell, T., Botvinick, M., Mar. 2009. Machine learn- ing classifiers and fMRI: A tutorial overview. NeuroImage 45 (1), S199–S209. Poldrack, R. A., Halchenko, Y. O., Hanson, S. J., Nov. 2009. De- coding the Large-Scale Structure of Brain Function by Clas- sifying Mental States Across Individuals. Psychological Sci- ence 20 (11), 1364–1372. Serences, J. T., Saproo, S., Mar 2012. Computational advances towards linking bold and behavior. Neuropsychologia 50 (4), 435–46. Soon, C. S., Brass, M., Heinze, H.-J., Haynes, J.-D., Apr. 2008. Unconscious determinants of free decisions in the human brain. Nature Neuroscience 11 (5), 543–545. Stokes, M., Saraiva, A., Rohenkohl, G., Nobre, A. C., Jun 2011. Imagery for shapes activates position-invariant representa-
1512.08339
1
1512
2015-12-28T08:00:27
Detecting local processing unit in drosophila brain by using network theory
[ "q-bio.NC", "physics.bio-ph" ]
Community detection method in network theory was applied to the neuron network constructed from the image overlapping between neuron pairs to detect the Local Processing Unit (LPU) automatically in Drosophila brain. 26 communities consistent with the known LPUs, and 13 subdivisions were found. Besides, 45 tracts were detected and could be discriminated from the LPUs by analyzing the distribution of participation coefficient P. Furthermore, layer structures in fan-shaped body (FB) were observed which coincided with the images shot by the optical devices, and a total of 13 communities were proven closely related to FB. The method proposed in this work was proven effective to identify the LPU structure in Drosophila brain irrespectively of any subjective aspect, and could be applied to the relevant areas extensively.
q-bio.NC
q-bio
Detecting local processing unit in drosophila brain by using network theory Dongmei Shi1,2, Chitin Shih3,4, Chungchuan Lo5 , Yenjen Lin1, and Annshyn Chiang1,6,7 Brain Research Center, National Tsing Hua University, Hsinchun 30013 Taiwan, Department of Physics, Bohai University, Jinzhou Lianning 121000, P.R.C, Department of Physics, Tunghai University, Taichung 40704 Taiwan, Physics Division, National Center for Theoretical Sciences, Hsinchu 30043, Taiwan, Institute of Systems Neuroscience, National Tsing Hua University, Hsinchu 30013, Taiwan, Institute of Biotechnology, National Tsing Hua University, Hsinchu 30013, Taiwan 1. 2. 3. 4. 5. 6. 7. Genomics Research Center, Academia Sinica, Taipei 11529, Taiwan Community detection method in network theory was applied to the neuron network constructed from the image overlapping between neuron pairs to detect the Local Processing Unit (LPU) automatically in Drosophila brain. 26 communities consistent with the known LPUs, and 13 subdivisions were found. Besides, 45 tracts were detected and could be discriminated from the LPUs by analyzing the distribution of participation coefficient P. Furthermore, layer structures in fan-shaped body (FB) were observed which coincided with the images shot by the optical devices, and a total of 13 communities were proven closely related to FB. The method proposed in this work was proven effective to identify the LPU structure in Drosophila brain irrespectively of any subjective aspect, and could be applied to the relevant areas extensively. Keywords: Drosophila Local processing unit Complex networks Modularity Instruction Some neuron circuits in Drosophila are believed worth studying in understanding the fundamental features of neurons by most neuroscientists [1-4]. Genetic screens in Drosophila have identified many genes involved in neural development and function [5-9]. One reason is the power of the Drosophila genetic toolbox [10-12], another basic reason is that fruit flies have about 1000-fold fewer neurons, while exhibit extensive cognitive abilities. Local processing unit (LPU) [13-14] was proposed in constructing a mesoscopic map of the brain circuit in Drosophila. A LPU is defined as a brain region consisting of its own local interneurons (LN) whose nerve fibers are completely restricted to this region. Further, each LPU is contacted by at least one nerve tract. LPU is a functional unit dealing with the signals, which plays a critical part in information transmission in the brain. Connectomics-Based analysis of information flow in the drosophila brain based on the LPUs has been studied recently, and it was shown that the network showed hierarchical structure, small world, and rich-club organization [14]. Normally, LPU and its boundary are determined by observation based on the optical devices, which exists some subjective factors. In this work, complex network theory [15-18] was introduced to detect the LPU structure in Drosophila brain based on LPU's definition. It was proven that LNs and neuron tracts could be both identified successfully, and so that LPU would be determined by LNs and the corresponding neuron tracts, irrespectively of any subjective aspect. Methods Community structure (Figure 1) [19]: A network is realized to have the community structure if the nodes of network can be easily grouped into sets of nodes such that each set of nodes is densely connected internally. Modularity Q: Modularity is one measure of the structure of networks or graphs. It was designed to measure the strength of division of a network into modules (also called groups, or communities). where , and is the module which the node i belongs to. is the strength of node i, and is the element in adjacency matrix. is determined by maximizing the modularity Q, and larger Q indicates more distinct community structure in the networks. Participation coefficient P: where is the strength of node i connected to the nodes in module , and is the total strength of node i in the network. is close to 0 if most of the edges node i linked to are in its own module . On the other hand, is close to 1 if the links are uniformly distributed among all the modules of the network. Simulations Neuron network construction: Neuron network was based on the 23380 female fruit fly data. The nodes in the network indicated the neurons of brain, and the links represented the interactions among the neurons. The corresponding adjacency matrix was constructed as follows: the matrix dimension was the total number of neurons, and the matrix element was the overlapping strength in space of neuron i and neuron j. The neuron network was an undirected-weighted complex network. Relations between LPU in the brain neuron system and community structure in a network (Figure 2): Neurons in the fruit fly brain could be categorized into two functional distinct populations: local interneurons (LNs) whose processes were restricted within a single brain region, and projection neurons (PNs) whose dendrites and axons connected two or more brain regions. Local Processing Unit (LPU) was defined as a brain region consisting of its own LNs population whose nerve fibers were completely restricted to the region, and meanwhile, each LPU was connected by at least one neural tract. So LPU boundary could be determined by LNs and the neuron tract fibers restrained to this region. On other hand, according to the definition of community structure in a network, two types of nodes could be classified: nodes whose links were all in a single community (P=0), and nodes whose links were distributed into several communities (P>0). Accordingly, a correspondence could be built up between LPU in the brain and community structure in a network: the LNs corresponded to the nodes with P=0, and PNs corresponded to the nodes with P>0. Figure 2 showed the representation of this relation. Community division: Based on the analysis above, detecting LPU structure was tantamount to find or form the communities in which participation coefficient P for all the nodes met P=0. Firstly, modularity calculation was used to divide the neuron networks into communities. In order to detect the basic LPU structure in which no smaller LPUs existed, modularity calculation was performed three times sequentially based on the original neuron network. Precisely (Fig. 3), the original network was divided into 8 communities by calculating Q (Cal. Q), and these 8 communities were called level 1 communities (L1-Comm). Then, L1-Comm would be further divided into smaller communities which were called level 2 communities (L2-Comm), and further, L2-Comm would be done by the same analogy into level 3 communities (L3-Comm). Taking the 5th community on level 1 for example, it was seen in Fig. 3 that the 5th community on L1-Comm was divided into 3 small communities on level 2, 5-1, 5-2, 5-3. And the second community 5-2 on level 2 could be further divided into 3 smaller communities on level 3, 5-2-1, 5-2-2, 5-2-3. And so on, for each of the 8 communities on level 1. Community detection method: Representation of community detection method was shown in Fig. 4, one can see that three steps were needed to detect the LPU structure. On the first step, the original brain network was divided into L3-Comm communities by Cal. Q, among which LPU communities were included. On step 2, neuron tract communities would be discriminated. It had been proven according to the statistical results in this study that a community was the tract community if , or if there existed a distinct segmentation point or P>0.5 (f(P)~0) in the distribution of P, then the neurons corresponding to the last segmentation belonged to a neuron tract. is the minimum of participation coefficient in this community, and f(P) is the distribution function of P. See an example of tract distribution in Fig. 5. LPU communities were left by manipulations in step 2, but there still existed PNs in these communities. In order to extract LNs from LPU communities, LPU community would be optimized in step 3. Based on correspondence between LPU and community structure (Fig. 2), our solution was to remove the nodes with large P piece by piece from the LPU community by modularity calculating successively, until of this community. The community could be realized isolated when in our study. Before our further study, there were two points needed to be discussed. According to the theory in community detection method, the nodes with participation coefficient P=0 corresponded to the LNs whose fibers were all restricted to one region. However, considering the existence of warping area differences, when the neurons were put into the standard brain [20-21], the critical value of P could not reach 0 accurately. Actually P should be a function of warping area, and was different in different brain regions. On other hand, considering that some LPUs might be destroyed by the successively divisions in step 1, operations on step 3 would be applied directly to the communities in level 1 to compensate the destroyed LPUs. Results Community detection method was applied to the Drosophila brain neuron system by means of the above manipulations. Figure 6 showed the LNs of different brain regions in the standard brain from forward, middle and backward levels of the brain, and it was proven that these LNs belonged to 26 known LPUs: DLP (dlp), MB (mb), AL (al), VLP (vlp), MED (med), LOB (lob), nod, FB, EB, LH (lh), LOP (lop), PB, CMP (cmp). Two subdivisions in VLP (vlp), three subdivisions in FB, and three subdivisions in MB (mb) were also illustrated. Generally, 49 LPUs and hubs had been found, and the lost LPUs in this work were mainly due to the fewer LNs supplied in these LPUs. 45 neuron tracts were discriminated by the community detection method, and were shown in Fig. 7. 9 types of neuron tracts were proven connected AL, which were sighed 1-9 in Fig. 8. Subdivisions in FB and MB were shown in Fig. 8 (1) and Fig. 8 (2). Three subdivisions in FB were detected which had been not observed in experiment, but were proven to be consistent with the results only based on FB data by applying the community detection method. Three subdivisions ( , , ) in MB (mb) shown in Fig. 8 (2) coincided with the observations in experiment by optical devices. Four layers in FB were detected, and illustrated in Fig. 9. Only neurons in Fig. 9 (3) were the LNs, and neurons in Fig. 9 (1), (2), (4) were all PNs according to the community detection method. Figure 10 shows the communities related to FB, including three LPU communities and other 10 tract communities. Apparently FB played an important part in information transmission in the brain. Summary We applied the community detection method to detect the LPUs in Drosophila brain system, and had proven that 26 known LPUs were detected, and subdivisions were also found. Precisely, three subdivisions in MB (mb), and two subdivisions in VLP (vlp) were consistent with the observations in experiments, and unknown three subdivisions in FB would give us important information to understand FB's internal structure. It was further proven that the same subdivisions in FB were also found if the community method was only applied to FB neuron system. Besides, layer structures in FB were detected, and one layer was a subdivision in FB, and other three layers were tract communities. A total of 13 communities were closely related to FB, including subdivisions of FB, neuron tract communities inside FB, and neuron tract communities on FB's boundary. Apparently, FB played an important part in information transmissions in Drosophila brain. Moreover, it had been proven that neuron tracts could be discriminated by analyzing the distributions of participation coefficient P. Community detection method was proven effective to detect LPU in Drosophila brain system in this study, and its boundary was determined by LNs and the neuron tracts totally, irrespectively of any subjective factor. This method would be more perfect if the participation coefficient P as a function of warping area was known, and would be applied to the relative areas extensively. Appendix [Detecting subdivisions in FB and subdivisions in neuron tract communities] Manipulations in step 3 of community detection method (Fig. 4) were directly applied to FB neuron system in which 544 neurons in FB had been obtained by experiment. Precisely, nodes with large P would be removed piece by piece from the FB neuron network successively until of this community. It had been proven that three subdivisions were found in FB coinciding with the results in the above simulations. Moreover, this method was also applied to neuron tract communities discriminated in step 2 in Fig. 4 in order to find the basic neuron tracts in these communities. Acknowledgments This work was supported by the MOE 5-Year-50-Billion Project and NSC Nano National Project in Taiwan, Specialized Foundation for Theoretical Physics of China (Grant No. 11247239), National Natural Science Foundation of China (Grants No. 11305017). [1]Shawn R. Olsen and Rachel L. Wilson, Cracking neural circuits in a tiny brain: new approaches for understanding the neural circuitry of Drosophila, Cell: Trends in Neurosciences, 31 (2008). [2]T.C. Holmes, etal, Circuit-breaking and behavioral analysis by molecular genetic manipulation of neural activity in Drosophila. In Invertebrate Neurobiology (North, G. and Greenspan, R.j., eds) Cold Spring Harbor Laboratory, (2007), pp. 19--52. [3]L. Luo etal, Genetic dissection of neural circuits, Neuron, 57 (2008), pp. 634--660. [4]S. Tang, etal, Visual pattern recognition in Drosophila is invariant for retinal position, Science, 305 (2004), pp. 1020--1022. [5]G. Liu, etal, Distinct memory traces for two visual features in the Drosophila brain, Nature, 439 (2006), pp. 551--556. [6]K. Neuser, etal, Analysis of a spatial orientation memory in Drosophila, Nature, 453 (2008), pp. 1244--1247. [7]C. Faucher, etal, Behavioral responses of Drosophila to biogenic levels of carbon dioxide depend on lif-stage, sex and olfactory context, J. Exp. Biol. 209 (2006), pp. 2739-2748. [8]A. Yurkovic, etal, Learning and memory associated with aggression in Drosophila melanogaster, Proc. Natl. Acad. Sci. U. S. A., 103 (2006), pp. 17519--17524. [9]J.W. Bohland, C. Wu, H. Barbas, H. Bokil, M. Bota, H.C. Breiter, H.T. Cline, J.C. Doyle, P.J. Freed, R.J. Greenspan, etal, A proposal for a coordinated effort for the determination of brain-wide neuroanatomical connectivity in model organisms at a mesoscopic scale, PLOS Comput. Biol. 5 (2009), e1000334. [10]S.R. Olsen, and R.I. Wilson, Cracking neural circuits in a tiny brain: New approaches for understanding the neural circuitry of Drosophila, Trends Neurosci. 31 (2008), pp. 512--520. [11] A. Claridge-Chang, R.D. Roorda, E. Vrontou, L. Sjulson, H. Li, J. Hirsh, and G. Miesenböck, Writing memories with light-addressable reinforcement circuitry. Cell 139 (2009), pp. 405--415. [12]M.J. Krashes, S. DasGupta, A. Vreede, B. White, J.D. Armstrong, and S. Waddell, A neural circuit mechanism integrating motivational state with memory expression in Drosophila. Cell 139 (2009), pp. 416--427. [13] A.S. Chiang, C.Y. Lin, C.C. Chuang, H.M. Chang, C.H. Hsieh, C.W. Yeh, C.T. Shih, J.J. Wu, etal, Three-Dimensional Reconstruction of Brain-wide Wiring Networks in Drosophila at Single-Cell Resolution}, Current Biology 21 (2011), pp. 1--11. [14]C.T. Shih, O. Sporns, S.L. Yuan, T.S. Su, Y.J. Lin, C.C. Chuang, T.Y. Wang, C.C. Lo, R.J. Greenspan, A.S.Chiang, Connectomics-Based Analysis of Information Flow in the Drosophila Brain, Current Biology 25 (2015), pp. 1249--1258. [15] N. Ganguly, A. Deutsch and A. Mukherjee, Applications Science, 978-0-8176-4750-6. [16] A. Barrat, M. Barthelemy and A. Vespignani, Dynamical processes on complex networks, Cambridge University Press, (2008), ISBN 978-0-521-87950-7. [17]G. Caldarelli, Scale-Free Networks, Oxford University Press, (2007), ISBN 978-0-19-921151-7. [18]E. Estrada, The Structure of Complex Networks: Theory and Applications, Oxford University Press, (2011), ISBN 978-0-199-59175-6. [19] M.E.J. Newman, Modularity and community structure in networks, Proc. Natl. Acad. Sci. USA 103 (2006), pp. 8577--8582. [20]A.W. Toga, Neuroimage databases: The good, the bad and the ugly, Nat. Rev. Neurosci. 3 (2002), pp. 302--309. [21]K. Rein, M. Zöckler, M.T. Mader, C. Grübel, and M. Heisenberg, The Drosophila standard brain, Curr. Biol. 12 (2002), pp. 227--231. Dynamics On and Of Complex Networks, ISBN to Biology, Computer and the Social Sciences, (2009) Figure 10. 13 communities closely related to FB. Figure 9. Layer structures in FB. Figure 8. (1) Subdivisions in FB; (2) Subdivisions in MB (mb). Figure7. 45 neuron tracts in the standard brain, and neuron tracts signed 1-9 were the tracts connecting the AL (al). Figure6. LNs in different regions in the standard brain from forward, middle and backward level, respectively. Figure5. Tract community discrimination. Figure 4. Graphic representation of community detection method. Step 1, community division; step 2, neuron tract discrimination; step 3, LPU community optimization. Figure3. Community division by calculating modularity ($Calc. Q$). The original brain neuron system was divided into 8 communities called level 1 community ($L1-Comm$), and these 8 communities could be further divided into sub-communities called level 2 community ($L2-Comm$). Based on $L2-Comm$, they would be further divided to communities called level 3 community ($L3-Comm$). Taking the 5th community on level 1 for example, it was divided into 3 smaller communities 5-1,5-2,5-3 on level 2, and community 5-2 was further divided into 3 sub-communities on level 3. Figure 2. Relations between LPU and community. (1) Black-solid curves represented the local interneurons (LNs), and other color-solid curves indicated the projection neurons (PNs). Green, blue, and red-solid curves represented neuron tracts; (2) Nodes in the same shadow formed a community, and two types of nodes were included, type of node A whose links were all in a single community ($P_{A}=0$), and type of B whose links were distributed into one more communities ($P_{B}>0$). Figure 1. Representation of community structure in a graph. Nodes in the same shadow form a community.
1309.1654
2
1309
2013-09-09T20:52:59
A Generic Approach to Solving Jump Diffusion Equations with Applications to Neural Populations
[ "q-bio.NC" ]
Diffusion processes have been applied with great success to model the dynamics of large populations throughout science, in particular biology. One advantage is that they bridge two different scales: the microscopic and the macroscopic one. Diffusion is a mathematical idealisation, however: it assumes vanishingly small state changes at the microscopic level. In real biological systems this is often not the case. The differential Chapman-Kolmogorov equation is more appropriate to model population dynamics that is not well described by drift and diffusion alone. Here, the method of characteristics is used to transform deterministic dynamics away and find a coordinate frame where this equation reduces to a Master equation. There is no longer a drift term, and solution methods there are insensitive to density gradients, making the method suitable for jump processes with arbitrary jump sizes. Moreover, its solution is universal: it no longer depends explicitly on the deterministic system. We demonstrate the technique on simple models of neuronal populations. Surprisingly, it is suitable for fast neural dynamics, even though in the new coordinate frame state space may expand rapidly towards infinity. We demonstrate universality: the method is applicable to any one dimensional neural model and show this on populations of leaky- and quadratic-integrate-and-fire neurons. In the diffusive limit, excellent approximations of Fokker-Planck equations are achieved. Nothing in the approach is particular to neuroscience, and the neural models are simple enough to serve as an example of dynamical systems that demonstrates the method.
q-bio.NC
q-bio
1 A Generic Approach to Solving Jump Diffusion Equations with Applications to Neural Populations Marc de Kamps1 1 School of Computing, University of Leeds, Leeds, West Yorkshire, United Kingdom ∗ E-mail: Corresponding [email protected] Abstract Diffusion processes have been applied with great success to model the dynamics of large populations throughout science, in particular biology. One advantage is that they bridge two different scales: the microscopic and the macroscopic one. Diffusion is a mathematical idealisation, however: it assumes vanishingly small state changes at the microscopic level. In real biological systems this is often not the case. The differential Chapman-Kolmogorov equation is more appropriate to model population dynamics that is not well described by drift and diffusion alone. Here, the method of characteristics is used to transform deterministic dynamics away and find a coordinate frame where this equation reduces to a Master equation. There is no longer a drift term, and solution methods there are insensitive to density gradients, making the method suitable for jump processes with arbitrary jump sizes. Moreover, its solution is universal: it no longer depends explicitly on the deterministic system. We demonstrate the technique on simple models of neuronal populations. Surprisingly, it is suitable for fast neural dynamics, even though in the new coordinate frame state space may expand rapidly towards infinity. We demonstrate universality: the method is applicable to any one dimensional neural model and show this on populations of leaky- and quadratic-integrate-and-fire neurons. In the diffusive limit, excellent approximations of Fokker-Planck equations are achieved. Nothing in the approach is particular to neuroscience, and the neural models are simple enough to serve as an example of dynamical systems that demonstrates the method. Author Summary Biological systems stand out by their complexity and multi-scale interactions. Modeling techniques that can explain the behavior of systems - the macroscopic level - in terms of their individual components - the microscopic level - are very valuable. Diffusion models do that, so, unsurprisingly, they pervade biology. One key assumption underlying the validity of the diffusion approach is that state changes of individuals are minuscule. In real biological systems this can often not be justified. The method presented here dispenses with the assumption and broadens the application range of population modelling considerably. Moreover, the method reduces the mathematically difficult problem of solving a partial differential equation to a simpler one: understanding the noise process that gave rise to it. We demonstrate the approach on neuronal populations. We envisage that the study of networks of such populations will elucidate brain function, as large-scale networks can be simulated with unprecedented realistic neural dynamics. This may help explain imaging signals, which after all are the observable collective signal of large groups of neurons, or it may help researchers to develop their ideas on neural coding. We also expect that the technique will find broad application outside of neuroscience and biology, in particular in finance. Introduction Diffusion models are of crucial importance to biological modeling; literally, text books have been filled on the subject, e.g. [1, 2]. The application domain is enormous: fluid dynamics, chemotaxis [3], animal 2 population movement [4, 5], neural populations e.g. [6 -- 17] are just a few examples (see [18] for a recent review and further references). Diffusion models allow an understanding of group behaviour in terms of that of its constituents, i.e. they link macroscopic and microscopic behaviour. In general diffusion models emerge as follows: microscopic deterministic laws describing individual behavior in the absence of interactions are determined. Interactions among individuals and between individuals and the outside world are described by a stochastic process ('noise'). Diffusion arises if one assumes that changes in the microscopic state as a consequence of an interaction are vanishingly small. Sometimes, this assump- tion is well justified: for example, in Brownian motion these interactions are due to the collisions of individual molecules, interactions that are indeed minuscule compared to the macroscopic state. Often, this assumption is not appropriate, but made in order to use the familiar mathematical machinery of diffusion equations. For example, in computational neuroscience the diffusion approximation hinges on the assumption that synaptic efficacies - a measure for how strongly neurons influence each other - are small. This is simply not the case in many brain areas, e.g. [19], and leads to sizable corrections to diffu- sion results [20]. Nevertheless, the diffusion approximation predominates in computational neuroscience. Given the widespread distribution of diffusion models in biology, it stands to reason that this situation is common. Experience from computational neuroscience [20] and finance - a field that has put considerable effort into understanding jump diffusion equations e.g. [21 -- 23] - has shown that results valid beyond the diffusion limit are very hard to obtain, limited to a specific study and often require a perturbation approach [20]. Due to their nature, jump diffusion processes may lead to very jagged, locally discontinuous probability density profiles [24]. For this reason numerical solution schemes that are well established for diffusion processes may be unsuitable for finite size jumps. Some solution schemes are applicable for finite jumps if it can be assumed that the underlying density profile is smooth. Brennan and Schwartz [25] demonstrated that an implicit scheme for solving the Black-Scholes equation is equivalent to a generalized jump process. The assumption must be justified, however. In contrast, the method presented here makes no assumptions about the structure of the underlying structure of the density profile at all. Below, we will present an example that can be easily modeled by the method presented here, but where we are not aware of more conventional discretization schemes that would apply. Standard text books such as [24, 26] demonstrate how deterministic laws of motion for individuals and a Master equation describing the noise process under consideration can be combined. This results in the so-called differential Chapman-Kolmogorov equation, a partial integro-differential equation in probability density (it is derived in the context of neuronal populations in "Methods: Derivation of the population density equation"). When the assumption is made that the transitions due to the stochastic processes induce very small state transitions, this equation reduces to a partial differential equation of the Fokker-Planck type [24, 26]. Diffusion models ultimately derive from this equation. Clearly, a solution method for the Chapman-Kolmogorov equation is highly valuable: it would still allow the study of diffusion processes, but also processes with large jumps. Below, we will give an example of the gradual breakdown of diffusion. This opens up an important application area: it will be possible to re-examine results already obtained in the diffusion limit and to investigate the consequences of finite jump sizes. Many expositions immediately proceed to the diffusion regime. Here we follow a radically different approach: we use the method of characteristics to define a coordinate frame that co-moves with the flow of the deterministic system. In this frame the differential Chapman-Kolmogorov equation reduces to the Master equation of the stochastic process under consideration. This approach neatly sidesteps the need for solving a partial (integro-) differential equation as the Master equation is a set of ordinary differential equations, and can be solved by relatively simple numerical methods. This simplicity comes at a price, however: the probability density is represented in an interval that is moving with respect to the new coordinate frame. How this interval behaves becomes dependent on the topology of the flow of the deterministic system. We will illustrate the problem on two neuron models 3 that represent two ends of the dynamic spectrum. Leaky-integrate-and-fire neurons have slow underlying deterministic dynamics. Quadratic-integrate-and-fire neurons model the rapid onset of a spike. The dynamics is so fast that the original interval representing the probability density profile moves to infinity in finite time in the new coordinate frame. We find that these neural models are a demonstration of the broad applicability of the technique. Not only are they simple examples of dynamical systems that can easily be understood without a background in computational neuroscience. They also provide boundary conditions that any good solution method must be able to handle: absorbing boundaries. In general neurons spike and then return to pre-spike conditions. In one dimensional neural models this is simulated by resetting the neuronal state as soon as an absorbing boundary (the 'threshold') is reached. The reintroduction of a neuron once it has hit the threshold (a 'reset') is particular to neural systems. It is a marked difference from financial derivatives - as options become worthless after their expiration date - and one reason why results from finance do not carry over to neuroscience despite similarities. The method is manifestly stable, and the jump size is immaterial. The method can be considered a generalization of the diffusion approach: for a stochastic process characterized by small jumps diffusion results are recovered, as one would expect. However, it is not always practical to study diffusion by a small finite jump process as the convergence to diffusion results is not always uniform and a direct application of the Fokker-Planck equation may then give better results. We believe the main application of the method is for processes with truly finite jumps, or to study deviations from the diffusion limit. Results We obtained the following results. 1. The method of characteristics was employed to define a new coordinate frame that moves along the flow of deterministic dynamical system. In the new coordinate frame the differential Chapman- Kolmogorov equations transforms to the time dependent Master equation of the noise process. This result was used earlier in the limited setting of leaky-integrate-and-fire neurons [27,28]. It is restated here for convenience. 2. The characteristics themselves were used to define a representation of the probability density. This is a key new insight as it allows an accurate probability representation even when the deterministic process is fast. This is explained in some detail below. 3. The density representation is dependent on the topology of the deterministic flow. Although the number of topologies is typically limited, potentially this could require a novel implementation of the algorithm. Somewhat surprisingly, we found that the simplest implementation of the density is for a deterministic process that corresponds to periodically spiking neurons. Importantly, we found that under broad conditions it is possible to 'switch topology' by a process we dubbed 'current compensation'. It is possible to modify the topology of the deterministic neural dynamics by adding a DC current to the neurons. We compensate by renormalizing the noise input such that its mean is lower by an amount equal to the DC current. Therefore, often a single topology suffices. 4. The algorithm that results from the application of the technique is independent of the neural model. Exploring different models does not require recoding of the algorithm, but merely the recalculation of a grid representation. The algorithm can be considered to be a universal population density solver for one dimensional neural models whenever current compensation is appropriate. We delivered an implementation of the algorithm in C++ that is publicly available (http://miind.sf.net). We used this implementation in the examples that illustrate the use of this algorithm. 4 Transforming away neural dynamics Assume that a neuron is characterised by a vector ~v ∈ M summarising its state, where M is a open subset of Rn. Further, assume that a smooth vector field ~F (~v) exists everywhere on M and that a density function ρ(~v, t) is defined for every ~v ∈ M . Now consider a large population of neurons. ρ(~v)d~v is the probability for a neuron to have its state vector in d~v. It is also assumed that the population is homogeneous in the sense that interactions between neurons and the outside world can be accounted for by the same stochastic process for all neurons (although individual neurons each see different realisations of this process). In the absence of noise, neurons follow trajectories through state space determined by the flow of vector field ~F (~v): τ d~v dt = ~F (~v), (1) where τ is the neuron's time constant. We must allow for the possibility of v being driven across ∂M , the edge of M , at which time it must be reset, immediately or after a refractive period at v = Vreset. As explained in "Methods: Derivation of the Population Density Equation", conventional balance arguments lead to an advection equation for the density. When a noise process is also taken into consideration, this equation becomes: ∂ρ ∂t + ∂ ∂~v · ( ~F ρ τ ) =ZM d ~w {W (~v ~w)ρ( ~w) − W ( ~w ~v)ρ(~v)} (2) W ( ~w ~v) is the probability per unit time for a noise event that will cause a state transition from d~v to d ~w. This equation is known as the differential Chapman-Kolmogorov equation [24]. Under the usual assumptions that guarantee the existence and uniqueness of solutions of Eq. 1 on M , one can find integral curves ~v′(t, ~v0(t0)). If ~v0 and t0 are suitably chosen on ∂M , these curves cover the entire manifold M and every point of M is uniquely defined by a coordinate pair (~v′, t). These curves are the characteristics of Eq. 2. Applying the method of characteristics one finds that the total time derivative of the density in the (~v′, t) system is given by: dρ′ dt =ZM′ where ρ′(~v′, t) = eR t ∂F (~v′ ) τ ∂~v′ dt′ ρ(~v′, t). d ~w′nW (~v′ ~w′)ρ′( ~w′) − W ( ~w′ ~v′)ρ′(~v′)o (3) For simplicity, the remainder of the paper will consider one dimensional neuronal models subject to external Poisson distributed spike trains, each spike causing an instantaneous jump in the membrane potential (the case of more than one input, or step sizes h that originate from a distribution of synaptic efficacies can be handled easily): leading to: W (v′ v) = νδ(v′ − h − v), dρ dt = ν {ρ(v′ − v′h, t) − ρ(v′, t)} + r(t)δ(v′ − v′reset), (4) (5) Here r(t) =R v′th ρ(w, t)dw; the δ peak reflects the reset of neurons that spiked to their reset potential. Equation 5 is the Master equation of an inhomogeneous Poisson process. When considered over a sufficiently short time scale, the solution of this process can be approximated by a homogeneous one, for which efficient numerical and analytic methods are available [28]. The main technical problem associated with solving Eq. 5 is to find a grid that adequately represents the density, not only at t = 0, but also at later times. Note that in v′-space all evolution is due to synaptic input, as in the absence of noise dρ(v′,t) v′th,h dt = 0. 5 10 ) s m ( t 8 6 4 2 10 ) s m ( t 8 6 4 2 0 10 5 5 0 10 V(t=0) (mV) 15 20 0 10 5 10 15 20 25 30 5 0 V(t=0) (mV) Figure 1. Solutions to Eq. 1 for LIF neurons (Vth = 20 mV, τ = 10 ms) for I = 0 mV (left) or I = 30 mV (right). This idea works well for leaky-integrate-and-fire neurons [27]: τ dV dt = −V + I, (6) where V is the membrane potential and I an external (non-stochastic) current. Here v′-space slowly expands, which can be accommodated for by adding points to the grid representing the density profile during simulation. Occasionally one must rebin to curb the resulting growth of the grid (see [27] for details). In general neuronal dynamics is not slow compared to synaptic input, e.g. consider the quadratic- integrate-and-fire model given by: τ dV dt = V 2 + I, (7) where V is a rescaled dimensionless variable. The characteristics can be calculated analytically (Table 1), and are shown in Fig. 2. Their topology depends on the value of I. It is clear that some curves run away to infinity in finite time: these neurons are intrinsically spiking. Upon reaching infinity, or some threshold potential Vth, these neurons are reintroduced in the system at V = Vreset. At first sight, it seems hard to find an adequate density representation without being forced to rapid and expensive rebinning operations that must be applied in brief succession. However, by adopting a grid whose bin limits are defined by the characteristics rebinning operations can be avoided altogether. Fast dynamics will result in large bins, slow dynamics in small ones. Large bins, however, do not introduce 10 ) s m ( t 8 6 4 2 0 10 0 5 15 V(t=0) (rescaled) 10 5 10 ) s m ( t 8 6 4 2 20 0 10 0 5 15 V(t=0) (rescaled) 10 5 10 ) s m ( t 8 6 4 2 20 0 10 0 5 15 V(t=0) (rescaled) 10 5 6 20 Figure 2. Characteristics for the values of I: I = −10 (top), I = 0 (middle) and I = 10 (bottom), on an interval D = [−10, 10]. I > 0 I = 0 I < 0 t = τ t = τ √I narctan V V0 − 1 lnn( V − t = τ 2√I √I V +√I V √I − arctan V0√Io V (t) = √I tann√I t V (t) = V0 1−V0 V (t) = √I (√I+V0)e−2√I t (√I+V0)e−2√I t )( V0+√I √I t τ V0− )o τ + arctan( V0√I )o τ −(√I−V0) τ +(√I−V0) The solutions to Eq. 1, for different values of I. Table 1. QIF Characteristics inaccuracy, because all density within the same bin will share a common fate. This is true even after reset. This observation is crucial and applies quite generally. For this reason we will explain it in some detail in the next section. 1 7 Figure 3. Modelling advection requires only a pointer update. Top: At time t each probability bin is associated with a density interval. Bottom: Update at t + 1. Representing fast dynamics Starting with a model of the form of Eq. 1, we assume that dV more general cases later on. dt > 0 everywhere on (Vmin, Vth), considering Consider now the initial value problem, posed by Equation 1 with boundary condition V (t = 0) = V0. Assuming V0 = Vmin, the neuron will reach threshold Vth in finite time tperiod. The neuron will then be reintroduced at Vreset. For simplicity Vreset = Vmin will be assumed, for a practical implementation of the algorithm this is immaterial. Consider now a division of time interval tperiod into N time steps: tstep ≡ tperiod N (8) The density profile ρ(v) for given time t will be represented in a grid consisting of N bins. We define the bin limits as follows: each bin i (i = 0,··· , N − 1) corresponds to a potential interval [vi, vi+1) in the following way: v0 = Vmin, and when V (t = 0) = Vmin, vi is given by: vi ≡ V (itstep), i = 0,··· , N, (9) where V (t) is the solution of Eq. 1. Note that there are N + 1 bin limits and that by definition VN ≡ Vth. In general, the binning is non-equidistant. When a neuron spikes, for example, it traverses a considerable potential difference in a short period of time and the bins covering this traversal will be very large. The evolution of a population density profile defined on (Vmin, Vth) can now be modeled as follows. A probability grid P is created of size N . An array V contains the bin limits of P. We define element i of V, Vi ≡ vi. Assume an initial density profile ρ(v, 0) is defined at t = 0. We represent this profile by setting Pi to: (10) Pi ≡ ρ(vi, 0)(vi+1 − vi), 1The reset is a crucial difference between neuroscience and finance. It is one reason why analytic results for financial derivatives do not carry over to neuroscience and why analytic results there are almost unobtainable. The method must be able to handle reset, but will also work for systems where individuals are not reintroduced. 8 so that Pi approximates the total probability between V = vi and V = vi+1. Remarkably, having defined the contents of V and P, they can remain constant, yet describe advection: the evolution of the density in the absence of synaptic input. To see this, consider the evolution over a period of time tstep. All neurons that previously had a potential between V = vi and V = vi+1, will now have a potential between V = vi+1 and V = vi+2, except for those that had a potential between V = vN−1 and Vth who now have a potential between V = Vmin and V = v1. So, the relationship between V and P changes, but not the actual array contents themselves. Specifically, the density profile at simulation time tsim = jtstep is given by: ρ(vi, tsim) = P(i−j) mod N vi+1 − vi This is all that is required to represent the density profile in v′-space; the computational overhead is negligible as an implementation of this idea only requires a pointer update without any need for shuffling data around (see Fig. 3). Equation 5 can be solved by assuming that during tstep neurons only leave their bin due synaptic input. It is also assumed that, although the magnitude of h is dependent on v′, it is constant during the small time tstep. The solution method is not really different from that described in [27 -- 29], but is complicated by the non-equidistance of the probability grid P. Where in the original implementation only the magnitude of the synaptic efficacy had to be recalculated in v′-space to find the transition matrix of the Poisson process, here a search is needed to locate the density bin that will receive probability from a given density bin, as the jump size h′ becomes dependent on v′ in v′-space. This issue is rather technical and is be explained in detail in "Methods: Synaptic Input: Solving the Master Equation". Switching Topologies: Current Compensation The idea is most easily illustrated for a Poisson process. In one dimension, for a single synaptic input with postsynaptic efficacy h receiving a single Poisson distributed input rate ν, the population density equation is given by [8]: ∂ρ ∂t + ∂ ∂v ( F (v)ρ(v) τ ) = ν {ρ(v − h) − ρ(v)} (12) Consider the limit h → 0, such that νh = constant. A Taylor expansion of the right hand side up to order h2 then gives: (11) (13) (14) (15) Now define: ∂ρ ∂t + ∂ ∂v (cid:26) F (v)ρ τ + ρhν − νh2 2 ∂ρ ∂v(cid:27) = 0 and Eq. 13 becomes a Fokker-Planck equation: µ ≡ νhτ σ2 ≡ νh2τ ∂ρ ∂t + ∂ ∂v (cid:26)ρ F (v) + µ τ σ2 2τ − ∂ρ ∂v(cid:27) = 0 This equation describes the evolution of the density due to the deterministic dynamics, determined by F (v) and an additive Gaussian white noise with parameters µ and σ. It is clear that Eq. 15 is invariant under the transformation: (cid:26) F (v) → F (v) + Ic µ → µ − Ic (16) 9 1 -110 -210 -10 1 -110 -210 -10 1 -110 -210 -10 0 10 V 0 10 V 0 10 V Figure 4. Density profile of QIF population decorrelated by large synapses. Line: population density; markers: Monte Carlo simulation. for any constant current value Ic, i.e. one add a DC component to a neuron when a similar mean contribution is subtracted from its input. This result is valid beyond the diffusion approximation, as it still holds when higher than second order derivatives are considered in the procedure used to derive Eq. 15. Table 2. Default simulation parameters for QIF neurons. N τ τref Vmin Vmax Vreset Ic σc hdif f 300 10−2 s 0 s -10 10 -10 0.2 0.01 0.03 r r r 10 Examples We will give a number of examples. For quadratic-integrate-and-fire neurons the default simulation parameters are given Tab. 2. In the remainder only deviations will be listed. Example 1: quadratic-integrate-and-fire neurons with extremely large synapses Figure 4 demonstrates the validity of the method beyond the diffusion limit. A periodically firing group of neurons, firing in synchrony at the start of the simulation, is decorrelated by a low frequency (5 Hz) high impact (h = 5, this is half of the entire interval!) input. The density profile is a slowly collapsing delta-peak, travelling along the characteristics of a periodically firing quadratic-integrate-and-fire neuron for several seconds, before reaching its steady state profile. We show the density profile in Fig. 4 at t = 0.02, 0.12 and 9.9 s. At no moment the population density is well described by a diffusion process, and the steady state firing rate deviates considerably from that predicted by numerical calculations based on the diffusion approximation. At t = 0.12 s, the density profile shows a fine structure that is indeed borne out by large-scale Monte Carlo simulations. We are not aware of a method sensitive to the density gradient that could have modeled this accurately. 5 4.5 4 3.5 3 2.5 2 1.5 1 0.5 0 -10 6 5 4 3 2 1 0 -10 = -0.1 = 0.1 t 0.019 s 0.029 s steady state analytic -8 -6 -4 -2 0 2 4 6 8 = 0.3 = 0.1 t 0.008 s 0.028 s 0.051 s steady state analytic -8 -6 -4 -2 0 2 4 6 8 10 V 10 V 1.4 1.2 1 0.8 0.6 0.4 0.2 0 -10 1 0.8 0.6 0.4 0.2 0 -10 = -0.1 = 0.5 t 0.019 s 0.029 s steady state analytic -8 -6 -4 -2 0 2 4 6 8 10 V = 0.3 = 0.5 t 0.008 s 0.028 s 0.051 s steady state analytic 0 10 V Figure 5. Examples of the evolution of the density of a population of QIF neurons under the influence of Gaussian white noise input. Example 2: quadratic-integrate-and-fire neurons in the diffusion limit A single Poisson input, or a combination of one excitatory and one inhibitory Poisson input, can emulate r m s r m s r m s r m s 11 Figure 6. Steady state density profiles in response to Gaussian white noise. a Gaussian white noise (see "Methods: Emulation of Gaussian white noise"). We will study both the density transient density profiles and the resulting firing rate response of the population (a neuron reaching threshold will emit a spike and reset to reset potential. The population firing rate is fraction of neurons that in a time interval, divided by that time interval). In Fig. 5 the transient density profiles are shown for four different (µ, σ) combinations. The negative input, small noise distribution (µ = −0.1, σ = 0.1) peaks at the stable equilibrium point. Almost no neurons are pushed across the second, instable equilibrium point and this population does not fire in response to its input (see Fig. 7). For higher noise values, the peak is smeared and some neurons are pushed across the unstable, leading to an appreciable firing rate; a deterministic current with similar mean would evoke no response (see Fig. 7). A larger mean µ = 0.3 evokes a clear response, and the peak of the density is clearly driven across V = 0. Unlike in Example 1, this peak will soon collapse. It will approach its steady state distribution but demonstrate small oscillations that are clearly visible in the firing rate for a long time. At higher noise, the steady state is reached faster and the firing rates transients die out much faster (Fig. 7). Figures 6 and 7 provide a summary of the diffusion results for µ ∈ [−0.5, 0.5] and σ = 0.1, 0.2, 0.5, 0.7. Figure 7 shows the transient firing rates are shown and compared to Monte Carlo simulations (red markers). Figure 6 shows the density profiles for large time t = 9.9 s, when profiles should assume their steady state (grey line, but barely visible since Monte Carlo and analytic results agree and are overlaid). Markers show Monte Carlo results, red lines are calculated directly from the diffusion approximation, using the integration scheme from [15]. Figure 8 shows the so-called gain curves, the steady state firing rates that can be read of Fig. 7 in black markers. Monte Carlo results are indicated by red markers, lines 12 0.5 -0.1 -0.2 -0.4 =-0.5 =0.7 =0.5 =0.2 =0.1 50 45 40 35 30 25 20 15 10 5 0 50 45 40 35 30 25 20 15 10 5 50 0 45 40 35 30 25 20 15 10 5 50 0 45 40 35 30 25 20 15 10 5 -0.3 50 45 40 35 30 25 20 15 10 5 0 50 45 40 35 30 25 20 15 10 5 50 0 45 40 35 30 25 20 15 10 5 50 0 45 40 35 30 25 20 15 10 5 50 45 40 35 30 25 20 15 10 5 0 50 45 40 35 30 25 20 15 10 5 50 0 45 40 35 30 25 20 15 10 5 50 0 45 40 35 30 25 20 15 10 5 50 45 40 35 30 25 20 15 10 5 0 50 45 40 35 30 25 20 15 10 5 50 0 45 40 35 30 25 20 15 10 5 50 0 45 40 35 30 25 20 15 10 5 0.3 0.2 0.1 0 50 45 40 35 30 25 20 15 10 5 0 50 45 40 35 30 25 20 15 10 5 50 0 45 40 35 30 25 20 15 10 5 50 0 45 40 35 30 25 20 15 10 5 50 45 40 35 30 25 20 15 10 5 0 50 45 40 35 30 25 20 15 10 5 50 0 45 40 35 30 25 20 15 10 5 50 0 45 40 35 30 25 20 15 10 5 50 45 40 35 30 25 20 15 10 5 0 50 45 40 35 30 25 20 15 10 5 50 0 45 40 35 30 25 20 15 10 5 50 0 45 40 35 30 25 20 15 10 5 50 45 40 35 30 25 20 15 10 5 0 50 45 40 35 30 25 20 15 10 5 50 0 45 40 35 30 25 20 15 10 5 50 0 45 40 35 30 25 20 15 10 5 0.4 50 45 40 35 30 25 20 15 10 50 45 40 35 30 25 20 15 10 50 45 40 35 30 25 20 15 10 50 45 40 35 30 25 20 15 10 50 45 40 35 30 25 20 15 10 5 0 50 45 40 35 30 25 20 15 10 5 50 0 45 40 35 30 25 20 15 10 5 50 0 45 40 35 30 25 20 15 10 5 0 0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5 0 0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5 0 0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5 0 0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5 0 0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5 0 0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5 0 0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5 0 0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5 0 0 0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5 Figure 7. Transient firing rates inresponse to Gaussian white noise. 50 45 40 35 30 25 20 15 10 5 0 50 45 40 35 30 25 20 15 10 5 50 0 45 40 35 30 25 20 15 10 5 50 0 45 40 35 30 25 20 15 10 5 0 0 s m s s s 13 ) z H ( f 25 20 15 10 5 0 = 0.1 0.2 0.5 0.7 -0.4 -0.2 0 0.2 0.4 0.6 Figure 8. Gain curves. Lines: analytic steady state firing rates. Red error bars: Monte Carlo simulation, black markers: population density results. are calculated analytically from the diffusion approximation [30]. All results are in excellent agreement with each other. Example 3: The gain curve is changed by large synapses. A single Poisson input can only approximate Gaussian white noise as long as the variability is much smaller than the mean of the signal. Figure 9 (top) shows the efficacy h as a function of µ for different σ values, and hence indicate where deviations can be expected (namely for large h). Figure 9 shows a population receiving input from a single Poisson input with firing rate and efficacy calculated using Eq. 30. µ must not be interpreted as the DC contribution of a Gaussian white noise input. A value of I = −1(Ic = 1.1) was chosen and simulated both in Monte Carlo and using the population density approach. Here, one expects the population to fire when the neurons' negative I value is overcome, i.e. for µ > 1. The steady state values of the population density approach are in excellent agreement with the Monte Carlo, but both deviate from the gain curves that are calculated in the diffusion approximation (Fig. 9 (bottom)), precisely where this would have been expected according to Fig. 9 (top). Note, however, that for the first time it was necessary to smear the synaptic efficacies slightly (σ = 0.01). In the absence of smearing, there would have been a disagreement between Monte Carlo and data for low σ. This is due to the fact that the variability of the compensation current Ic cannot be chosen arbitrarily low as this would entail artificially high firing rates that would render the solution of the Master equation inefficient. A reasonable minimum value is σ = 0.01. As can be seen from Fig. 9, the jumps are really large for σ = 0.7 and low µ. Indeed, there the deviation from the diffusion approximation is substantial, but there is no visible disagreement between Monte Carlo and population density approach. This is all the more impressive because to model negative I values, current compensation with Ic = 1.1 had to be used, and demonstrates the viability of the current compensation approach. When the steady state density profile obtained with the population density method is compared to that obtained from the Monte Carlo, again good agreement is found, while a small but significant deviation from the density profile calculated from the steady state, of the Fokker-Planck equation Eq. 15 can be seen for large σ and low µ (Fig. 9 (top)). Example 4: Universality: quadratic- and leaky-integrate-and-fire neurons are handled by the same algorithm. m s 14 = = = = 0.1 0.1 0.1 0.1 0.2 0.2 0.2 0.2 0.5 0.5 0.5 0.5 0.7 0.7 0.7 0.7 h 1 0.8 0.6 0.4 0.2 0 0.5 0.6 0.7 0.8 0.9 1 1.1 1.2 1.3 1.4 1.5 ) z H ( f 25 20 15 10 5 0 = 0.1 0.2 0.5 0.7 0.6 0.8 1 1.2 1.4 Figure 9. For large synapses the gain curve changes. Top: Jump size for a single Poisson input emulating Gaussian white noise. Bottom: Monte Carlo (red markers), population density results (black markers) and diffusion prediction (lines). We use current compensation to turn leaky-integrate-and-fire neurons into spiking neurons. We add a small constant current Ic = 1.1 as in Eq. 6. Without this current the characteristics move away from threshold, with this current they cross threshold (compare Fig. 1 left and right). A Gaussian white noise current emulated by Poisson input with mean µ = −1.1 and small variability (σ = 0.05) is presented as input. This is combined with a single Poisson input of 800 Hz with synaptic efficacy h = 0.03. This example was used in earlier studies as a benchmark [8, 27]. In Fig. 10 the steady state density profile and the firing rate of the population are shown. They are compared to the simulation without current compensation, described earlier [27], using a density representation tailored to leaky-integrate-and-fire characteristics (red line). The results are almost identical. In the original simulation the steep peaks are the consequence of a single synaptic efficacy. Sirovich [29] has shown analytically that for the first peak the density profile is discontinuous, for the second peak the first derivative is discontinuous, and so on. This is not necessarily realistic in terms of neuroscience, but it demonstrates once more that the method can handle the jagged density profiles that result from finite size jump processes. In the current compensation version of the simulation, the small variability of the compensation current smears these peaks. The noise also slightly reduce the extremes of the firing rate (Fig. 10), but otherwise the agreement is excellent. The effect of the smearing due to the compensation current is equivalent to effects caused by spread in the synaptic efficacies [30]. Moreover, many brain systems receive an unspecific background that must be modeled as noise anyway, so current compensation can be m s s s s m s 15 2 1.8 1.6 1.4 1.2 1 0.8 0.6 0.4 0.2 0 20 18 16 14 12 10 8 6 4 2 0 0 i ) s / s e k p s ( e t a r g n i r i f 0 0.2 0.4 0.6 0.8 1 V (dimensionless) 0.05 0.1 0.15 0.2 0.25 0.3 t (s) Figure 10. Algorithm works for both QIF and LIF neurons. Top: steady state density profile for current compensation (black), dedicated LIF neuron implementation (red). Bottom: output firing rates. applied quite generally in neuroscience. Importantly, the only change that had to made to change over from quadratic- to leaky-integrate- and-fire neurons is in the calculation of array V, which is done during the initialization of the algorithm. Discussion We presented a method for numerically solving the differential Chapman-Kolmogorov equation for large jumps. We demonstrated the validity of the approach by comparing population density results for neu- ronal populations receiving input from large synapses to Monte Carlo simulations, and neurons receiving input from small synapses to analytic or numerical results obtained in a diffusion approach. We presented two examples (Example 1 and 4) that demonstrate that the method can handle arbitrary jump sizes and the extremely jagged density profiles that result. We also demonstrated that the algorithm is universal: it can handle any one dimensional neural, subject to the condition that a (pseudo) diffusive background of small variability is allowed. This point is obvious mathematically, the transformation from the differential Chapman-Kolmogorov equation to the Master equation of the noise process can always be performed [27]. The challenge is to find a suitable density representation. Periodically spiking neurons yield a particularly simple density representation, In practice the and using current compensation a non spiking neural model can be made to spike. r 16 Figure 11. High level overview of the approach. Neural model (red) only enters explicitly is in V. compensation current will always introduce some variability. Where variability is present already, for example because the input is noise or because synaptic efficacies are variable, this extra variability can be absorbed and has no bearing on the simulation results. The universality of the algorithm is reflected in the implementation. Although the neuronal model defines the grid boundaries of the density representation - the content of grid V, it does so before simulation starts. This means that neuronal models can be exchanged with great simplicity. To provide a neural model is to provide the contents of grid V. This is the only place were the neuronal model enters the method explicitly. This important point is emphasized in Fig. 11, which provides a high level overview of the method. We demonstrated the method using a Poisson process and showed that if the jump size becomes small, diffusion results are recovered. The use of Poisson statistics is appropriate in neural simulations. An individual neuron will receive at most of the order of 105 spikes per seconds. At such rates, a Poisson Master equation can solved efficiently. We found that 1 simulation second could be handled in between a 0.5 and 2.0 s real time, which is an order of magnitude faster than Monte Carlo - based on comparisons to leaky-integrate-and-fire 2 simulations in NEST [31] where a 10000 neuron population achieves real time ratios of approximately 30. We found an excellent recovery of diffusion results whenever the jump size is 5 % of the interval for low noise and of the order of 10 % for high noise conditions (Fig. 9). 2Quadratic-integrate-and-fire neurons were not implemented at the time of writing. We performed Monte Carlo simula- tions by writing an event-driven simulator based on the analytic solutions of Table 1. 17 It is conceivable that in other applications jump sizes are smaller. Artificially small jump sizes would lead to very high firing rates for a Poisson master equation. We find that adopting a jump size of less than 1 % of the interval leads to inconveniently high firing rates. The method is clearly best used for finite jump sizes. To study diffusion, it would be interesting to consider a finite difference scheme for solving Eq. 15 with F (v) = 0. This would allow a direct comparison to diffusion results, while still benefitting from the universality of this approach. For example, it would still be possible to compare diffusion in different neural models without the need to derive a separate Fokker-Planck equation for each case. Cox and Ross [32] have shown that explicit finite difference schemes for solving equations of the type 15 correspond to a jump process with three possibilities: a move left, right or no move in the grid. Brennan and Schwartz [25] have shown that implicit finite difference scheme corresponds to a jump process where jumps to every other position are allowed. For suitable parameters and in the limit of an infinitely fine grid both schemes converge to a diffusion process, where the implicit scheme is unconditionally stable. These results are interesting because they suggest suitable Master equations to arrive at good diffusion approximations. In the context of neural systems this discussion is to some extent academic as true diffusion conditions are not realized in the brain, but most analytic results on population densities have been derived in the diffusion approximation, and a comparison is valuable. The ideas presented here constitute an approach rather than an algorithm. There is an endless variation of noise processes and dynamical systems that can be combined. Colored noise can be handled by a two-step approach: the Master equation of a colored noise process can be simulated by leaky- integrate-and-fire population receiving white noise. The technique was demonstrated on one dimensional neural populations. Such models are simple enough to demonstrate the applicability of the technique to a wider audience. Whether it can be success- fully employed to more complex neural models is an open question. Fortunately, there is a tendency to move from complex conductance-based models to simpler low dimensional effective model neurons, see e.g. [33]. The population density approach still is competitive in two dimensions [14]. We hope that the technique will find widespread application. Depending on the application area, the methods described here may need adaptation. The current compensation technique may not work for all applications as the extra variability introduced by the compensation current might not always be acceptable. However, this is no fundamental objection to using this approach. Applying a combination of techniques described here and earlier [27, 28] useful density representations may still be found, although their implementation may be limited to a specific topology of the deterministic flow. Methods Solutions in absence of synaptic input Consider a neuron in the absence of synaptic input, but receiving a constant input current I. In general Equation 1 can be used to evolve the state of a neuron that has potential V0 at t = 0. For some cases analytic solutions exist. For leaky-integrate-and-fire neurons: V (t) = I − (I − V0)e− t τ (17) For quadratic-integrate-and-fire neurons, the solutions of Eq. 7 depend on the sign of I, and there are clear topological differences between solutions for different I. They are given in Tab. 1 and shown in Fig. 2. 18 Derivation of the population density equation The population density approach can be motivated as follows and is not restricted to one dimensional neuronal models. To reflect that, this section will consider the more general model: τ d~v dt = ~F (~v) (18) Assume that M is an open subset of Rn, where n is the dimensionality of the neuronal point model defined by Eq. 18. Assume that the neuronal state is given by point ~v in M. Further, assume that a density function ρ(~v, t) is defined on M and that ρ(~v, t)d~v represents the probability for a given neuron to be in the state space element d~v at time t. Consider the evolution of neuronal states for neurons in d~v during a short period of time dt. If external input is ignored, then under the influence of neuronal dynamics all neurons, and no others, that were in d~v at time t are in volume element d~v′ at time t′, with ~v′ and t′ defined by: Conservation of probability requires: (cid:26) t → t′ = t + dt ~v → ~v′ = ~v + ~dv = ~v + ~F τ dt J = so that, collecting all terms up to O(dt): dt +O(dt2), ∂~v ( 1 τ ∂ ~F (~v) ∂~v ρ + ∂ρ ∂t + ~F (~v) τ ∂ρ ∂~v) d~v = 0. · Since this equation must hold for infinitesimal but arbitrary volumes d~v, it follows that: ∂ρ ∂t + 1 τ ∂ ∂~v · ( ~F ρ) = 0 Equation 26 is an advection equation for probability flow under the influence of neuronal dynamics. When noise is also considered the equation becomes: ∂ρ ∂t + 1 τ ∂ ∂~v · ( ~F ρ) =ZM d ~w {W (~v ~w)ρ( ~w) − W ( ~w ~v)ρ(~v)} . (27) This is an example of what Gardiner calls the differential Chapman-Kolmogorov equation [24]. In the remainder only Poisson processes will be considered, but the method is applicable to any noise process that can be modeled by a Master equation. For a single Poisson input W (~v′ ~v) is given by: W (~v′ ~v) = νδ(v′0 − v0 − h) (28) Using: Equation 20 becomes: Since with: ρ(~v, t)d~v = ρ(~v′, t′)d~v′. ρ(~v′, t′) = ρ(~v, t) + dt ∂ρ(~v, t) ∂t + d~v · ∂ρ(~v, t) ∂~v , (cid:26)ρ(~v, t) + d~v · ∂ρ(~v, t) ∂~v + dt ∂ρ(~v, t) ∂t (cid:27) d~v′ = ρ(~v, t)d~v d~v′ = J d~~v, ∂ ~F (~v) ∂~v′ ∂~v = I + 1 τ (19) (20) (21) (22) (23) (24) (25) (26) 19 Here ν is the rate of the Poisson process, it is assumed that one of the neuronal state variables, v0 instantaneously transitions from v0 to v0 + h, where h is the synaptic efficacy. In a one dimensional neuronal model v0 is V , the membrane potential. Clearly, not all synapses in a population have a single value, and often one assumes: W (~v′ ~v) = νZ dhp(h)δ(v′0 − v0 − h), (29) where p(h) is often assumed to be Gaussian [8, 30]. In a network where many populations connect to any given population, multiple contributions of the form Eq. 29 must be included - one for each separate input. In principle, each input must be modeled separately, but often many inputs can be subsumed into a single one using the central limit theorem, leading to a Gaussian white noise input contribution. Sometimes kernels p(h) of separate inputs overlap and can be integrated into one input contribution. Figure 12. Emulation of a Gaussian white noise with Poisson inputs. Left: isoline of h = 0.01Vth, h = 0.1Vth. Right: synaptic efficacies are fixed at 3 % of Vth. Blue (red) lines are isolines of frequency of excitatory (inhibitory) inputs. 20 Figure 13. The Master equation follows from translating the contents of grid V the synaptic efficacy. Emulation of Gaussian white noise and the breakdown of the diffusion approx- imation It is evident that by choosing a small efficacy h and large input rate ν, the Poisson input approximates additive Gaussian white noise. These parameters must be chosen such that they provide the correct µ and σ. Inverting Eq. 14 gives: h = ν = σ2 µ µ2 τ σ2 (30) A single Poisson input can only emulate a Gaussian white noise when h is small compared to the relevant potential difference, e.g. for leaky-integrate-and-fire neurons when h ≪ Vth, or h ≪ I for quadratic- integrate-and-fire neurons, i.e. not for all values of (µ,σ) can values for h and ν be found so that the diffusion regime is reliably reproduced. This is borne out by Fig. 12 (left). Here a neuron is considered with τ = 10 ms, Vth(≡ θ) = 20 mV. For a (µ, σ) plane isolines of h are plotted: one where h is precisely at 1% of the threshold potential, one where it is at 10%. It is clear that a single Poisson input can emulate a broad range of µ values, but only at low σ. Using two inputs, one excitatory and one inhibitory, one can fix the synaptic efficacies at a low value compared to Vth. If this value is J one finds: µ = τ J ν(νe − νi) σ2 = τ νJ 2(νe + νi) (31) One then has two frequencies νe and νi as free parameters. This setup covers most of the (µ, σ) plane, except for values at low σ, where the input frequency of the inhibitory input falls below 0. In Fig. 12 only frequency isolines for positive input frequencies are shown. The boundary where the inhibitory input rate falls below 0 is forbidden for two inputs, but this is precisely the area covered by a single input (clearly if νi = 0, we have a single excitatory input). The arguments presented here cover a positive range of µ, similar arguments can be made for a net inhibitory input. In summary, one needs either one or two Poisson inputs to emulate a Gaussian white noise. In particular, this implies that we can use a single excitatory input as a model for the gradual breakdown of the diffusion limit: if we increase σ and calculate the synaptic efficacy using Eq. 30, we should see deviations from the diffusion limit as h can no longer be considered to be small. This is shown in Fig. 9 which shows the step size for a single Poisson input given µ, σ. For σ comparable to µ this step size is not necessarily small compared to the potential scale on which the neuronal model is defined and deviations of the diffusion approximation can be expected. 21 Whenever a Gaussian white noise with very small variability is required, this must be delivered by a single Poisson input. The firing rate is then inversely proportional to the jump size. For very low variability a numerical solution of the Poisson process may become inefficient. Synaptic input: solving the master equation Now consider a single Poisson input. The evolution of the density profile is straightforward in principle but complicated by the bins being non-equidistant. Formally the noise process can be represented as a transition matrix. Figure 13 how this matrix can be determined: the contents of grid V are shifted by an amount h, where h is the synaptic efficacy. Probability moves from potential V − h to V as consequence of an input spike. In large bins this means that neurons stay mostly within the same bin. Neurons that are in small bins may end up several bins higher. Below we will examine this process in detail. 10 5 0 5 10 0 10 20 30 40 0 10 20 30 40 0 10 20 30 40 0.0 0.2 0.4 0.6 0.8 0 10 20 30 40 Figure 14. Transition matrices for QIF (top) and LIF neurons (bottom). Loss of probability is indicated by black, gain by white. To determine the Poisson Master equation for excitatory input (h > 0), consider that every neuron arriving in the interval Di = [vi, vi+1) must have come from Di,h = [max(vi−h, Vmin), vi+1−h). Consider, the N + 1 integers defined by: pi ≡( vi − h < v0: vi − h ≥ v0: -1 arg min k vk ≤ vi − h < vk+1 (32) Now consider i = 0 ··· N − 1; if pi = pi+1, αi,pi is given by: vi+1 − vi vpi+1 − vpi αi,pi ≡ For pi = −1, αi,l = 0. If pi < pi+1, αi,l is given by: ; αil ≡  with: Finally, l = pi: pi < l < pi+1: l = pi+1: vpi +1−(vi−h) vpi +1−vpi θl 1 vi+1−h−vpi+1 vpi+1+1−vpi+1 θl ≡(cid:26) l < 0: l ≥ 0: 0 1 From this the Master equation at t = ktstep follows: 0 vpN +1−vN +h vpN +1−vpN 1. i < pN : i = pN : i > pN : βi ≡  = − Pi−k mod N + dPi−k mod N dt 22 (33) (34) (35) (36) (37) (38) (39) αi,lPl−k mod N pi+1 Xl=pi βlPl−k mod N ), N−1 +δiR( Xl=0 vk ≤ Vreset ≤ vk+1 R ≡ arg min k where R is defined by: and δ is the Kronecker δ. The term including it reflects the re-entry of neurons at the reset potential once they have spiked. For inhibitory input (h < 0), the formulae are almost identical as now v−h refers to a higher potential. Again, define pi ≡( vi − h > vN : N vi − h ≤ vN : arg min k vk ≤ vi − h < vk+1 Again consider i = 0 ··· N − 1; if pi = pi+1, αi, pi is given by: vi+1 − vi v pi+1 − v pi αi,pi ≡ ; if pi < pi+1, αi,l is given by: αil ≡  Finally, γi is defined as follows: l = pi: pi < l < pi+1: l = pi+1: v pi +1−(vi+h) v pi+1−v pi 1 θN−1− pi+1 vi+1+h−v pi+1 v pi+1+1−v pi+1 i < p0: i = p0: i > p0: 1 vp0+1−(v0+h) 0 vp0+1−vp0 γi ≡  (40) (41) (42) (43) The Master equation for an inhibitory input is: dPi−k mod N dt = −(1 − γi)Pi−k mod N + pi+1 Xl= pi αi,lPl−k mod N Note that Eqs. 38 and 44 are of the form: d ~P dt = M(k) ~P , 23 (44) (45) where M(e) is the transition matrix resulting from adding individual input contributions of the form 38 and 44. Solving the master equation Equation 38 and 44 were solved numerically using Runge-Kutta-Fehlberg integration. This is not optimal, but is very straightforward to implement, and yields a performance that is faster than Monte Carlo simulation by at least one order of magnitude. Note that in Eq. 38 one simply ignores the zero elements of M(e). The calculation of the time derivatives then scales as O(N )O(Ninput), where Ninput is the number of inputs that need to considered. In many neural systems, such as cortex it can be assumed that most inputs deliver a background that can be treated as a single noisy input [14, 30] Different neuronal models: different transition matrices Although the noise process and the neuronal dynamics are independent, the neuronal dynamics determines the representation of the density profile. The bin boundaries of the density profile in v-space, V are given by the evolution of the neuronal state, and therefore the transition matrix is directly dependent on the neuronal model. This is illustrated in Fig. 14, which shows the transition matrix for identical h, ν for quadratic-integrate-and-fire and leaky-integrate-and-fire neurons. First, it should be observed that to accommodate for spiking behaviour, the state space of the quadratic-integrate-and-fire neuron is much larger, for the leaky-integrate-and-fire neuron it is limited by the threshold, Vth. Second, the quadratic- integrate-and-fire neuron is a truly spiking model: once it spikes, its state runs away to infinity in finite time. The leaky-integrate-and-fire neuron runs over threshold towards the externally applied current in infinite time. It is clear that probability lost from a bin will end up in a higher bin, i.e. a bin corresponding to a higher potential, as expected for an excitatory event. For leaky-integrate-and-fire neurons the same is done. The differences between a spiking neuron model and a non-spiking model are clearly visible. The outer bins of the quadratic-integrate-and-fire grid are huge, because the potential difference bridged during a simulation step tstep is large during a spike. Relative to their size the outer bins are barely shifted with respect to each other, meaning that most probability originating from this bin during a synaptic event will end up there again. This makes intuitive sense: during a spike the influence of synaptic input is negligible. Indeed the transition matrix shows that no probability transfer takes place in the outer bins for quadratic-integrate-and-fire neurons. For leaky-integrate-and-fire neurons, the picture is different: neurons will mainly pass threshold due to synaptic input. After crossing threshold, their potential will be reset. The influence of the reset is clearly visible in the leaky-integrate-and-fire transition matrix as a horizontal line in the row corresponding to the reset bin. 24 Acknowledgments References 1. Ricciardi LM (1977) Diffusion Processes and Related Topics in Biology (Lecture Notes in Biomath- ematics). Springer-Verlag. 2. Murray JD (2007) Mathematical Biology: I. An Introduction (Interdisciplinary Applied Mathe- matics) (Pt. 1). Interdisciplinary applied mathematics. Springer, 3rd edition. 3. Alt W (1980) Biased random walk models for chemotaxis and related diffusion approximations. Journal of Mathematical Biology 9: 147 -- 177. 4. Skellam JG (1951) Random dispersal in theoretical populations. Biometrika 38: 147 -- 218. 5. Skellam J (1973) The Formulation and Interpretation of Mathematical Models of Diffusionary Processes in Population Biology. In: Bartlett MS, Hiorns RW, editors, The Mathematical Theory of the Dynamics of Biological Populations, New York: Academic Press. 6. Stein RB (1965) A Theoretical Analysis of Neuronal Variability. Biophysical Journal 5: 173 -- 194. 7. Knight BW (1972) The Relationship between the Firing Rate of a Single Neuron and the Level of Activity in a Population of Neurons. The Journal of General Physiology 59: 767 -- 778. 8. Omurtag A, Knight BW, Sirovich L (2000) On the simulation of large populations of neurons. Journal of Computational Neuroscience 8: 51 -- 63. 9. Brunel N, Chance F, Fourcaud N, Abbott L (2001) Effects of synaptic noise and filtering on the frequency response of spiking neurons. Physical Review Letters 86: 2186 -- 2189. 10. Haskell E, Nykamp DQ, Tranchina D (2001) Population density methods for large-scale modelling of neuronal networks with realistic synaptic kinetics: cutting the dimension down to size. Network 12: 141 -- 174. 11. Casti ARR, Omurtag A, Sornborger A, Kaplan E, Knight B, et al. (2002) A Population Study of Integrate-and-Fire-or-Burst Neurons. Neural Computation 14: 957 -- 986. 12. Fourcaud-Trocm´e N, Hansel D, van Vreeswijk C, Brunel N (2003) How spike generation mechanisms determine the neuronal response to fluctuating inputs. The Journal of Neuroscience 23: 11628 -- 11640. 13. Mattia M, Del Giudice P (2004) Finite-size dynamics of inhibitory and excitatory interacting spiking neurons. Physical Review E 70: 052903+. 14. Apfaltrer F, Ly C, Tranchina D (2006) Population density methods for stochastic neurons with realistic synaptic kinetics: Firing rate dynamics and fast computational methods. Network 17: 373 -- 418. 15. Richardson MJE (2007) Firing-rate response of linear and nonlinear integrate-and-fire neurons to modulated current-based and conductance-based synaptic drive. Physical Review E 76: 021919+. 16. Helias M, Deger M, Rotter S, Diesmann M (2011) Finite Post Synaptic Potentials Cause a Fast Neuronal Response. Frontiers in Neuroscience 5. 17. Dumont G, Henry J (2013) Population density models of integrate-and-fire neurons with jumps: well-posedness. Journal of Mathematical Biology 67: 453 -- 481. 25 18. Codling EA, Plank MJ, Benhamou S (2008) Random walk models in biology. Journal of The Royal Society Interface 5: 813 -- 834. 19. Sirovich L (2007) Populations of Tightly Coupled Neurons: The RGC/LGN System. Neural Com- putation 20: 1179 -- 1210. 20. Helias M, Deger M, Rotter S, Diesmann M (2010) Instantaneous Non-Linear Processing by Pulse- Coupled Threshold Units. PLoS Comput Biol 6: e1000929+. 21. Merton R (1976) Option pricing when underlying stock returns are discontinuous. Journal of Financial Economics 3: 125 -- 144. 22. Kou SG (2002) A Jump-Diffusion Model for Option Pricing. Management Science 48: 1086 -- 1101. 23. Cont R, Voltchkova E (2005) A Finite Difference Scheme for Option Pricing in Jump Diffusion and Exponential L´evy Models. SIAM Journal on Numerical Analysis 43: 1596 -- 1626. 24. Gardiner CW (1997) Handbook of Stochastic Methods. Springer, 2 edition. 25. Brennan MJ, Schwartz ES (1978) Finite Difference Methods and Jump Processes Arising in the Pricing of Contingent Claims: A Synthesis. The Journal of Financial and Quantitative Analysis 13: 461+. 26. Van Kampen NG (1992) Stochastic Processes in Physics and Chemistry, Volume 1, Second Edition (North-Holland Personal Library). North Holland, 2 edition. 27. de Kamps M (2003) A simple and stable numerical solution for the population density equation. Neural Computation 15: 2129 -- 2146. 28. de Kamps M (2006) An analytic solution of the reentrant Poisson master equation and its appli- cation in the simulation of large groups of spiking neurons. In: International Joint Conference on Neural Networks, 2006. pp. 102 -- 109. 29. Sirovich L (2003) Dynamics of neuronal populations: eigenfunction theory; some solvable cases. Network (Bristol, England) 14: 249 -- 272. 30. Amit DJ, Brunel N (1997) Model of global spontaneous activity and local structured activity during delay periods in the cerebral cortex. Cerebral cortex (New York, NY : 1991) 7: 237 -- 252. 31. Gewaltig MO, Diesmann M (2007) Nest (neural simulation tool). Scholarpedia 2: 1430. 32. Cox JC, Ross SA (1976) The valuation of options for alternative stochastic processes. Journal of Financial Economics 3: 145 -- 166. 33. Brette R, Gerstner W (2005) Adaptive exponential integrate-and-fire model as an effective descrip- tion of neuronal activity. Journal of Neurophysiology 94: 3637 -- 3642.
1706.05702
1
1706
2017-06-18T19:13:26
Perfect spike detection via time reversal
[ "q-bio.NC", "math.DG", "physics.bio-ph", "q-bio.QM" ]
Spiking neuronal networks are usually simulated with three main simulation schemes: the classical time-driven and event-driven schemes, and the more recent hybrid scheme. All three schemes evolve the state of a neuron through a series of checkpoints: equally spaced in the first scheme and determined neuron-wise by spike events in the latter two. The time-driven and the hybrid scheme determine whether the membrane potential of a neuron crosses a threshold at the end of of the time interval between consecutive checkpoints. Threshold crossing can, however, occur within the interval even if this test is negative. Spikes can therefore be missed. The present work derives, implements, and benchmarks a method for perfect retrospective spike detection. This method can be applied to neuron models with affine or linear subthreshold dynamics. The idea behind the method is to propagate the threshold with a time-inverted dynamics, testing whether the threshold crosses the neuron state to be evolved, rather than vice versa. Algebraically this translates into a set of inequalities necessary and sufficient for threshold crossing. This test is slower than the imperfect one, but faster than an alternative perfect tests based on bisection or root-finding methods. Comparison confirms earlier results that the imperfect test rarely misses spikes (less than a fraction $1/10^8$ of missed spikes) in biologically relevant settings. This study offers an alternative geometric point of view on neuronal dynamics.
q-bio.NC
q-bio
Perfect spike detection via time reversal J. Krishnan1 2 4, P.G.L. Porta Mana1, M. Helias1 2 3, M. Diesmann1 2, E. Di Napoli2 4 1Institute of Neuroscience and Medicine (INM-6), Jülich Research Centre, Germany 2Institute for Advanced Simulation (IAS), Jülich Research Centre, Germany 3Department of Physics, Faculty 1, RWTH Aachen University, Aachen, Germany 4Aachen Institute for advanced study in Computational Engineering Science (AICES), RWTH Aachen University, Germany Correspondence*: Jeyashree Krishnan [email protected] ABSTRACT Spiking neuronal networks are usually simulated with three main simulation schemes: the classical time-driven and event-driven schemes, and the more recent hybrid scheme. All three schemes evolve the state of a neuron through a series of checkpoints: equally spaced in the first scheme and determined neuron-wise by spike events in the latter two. The time-driven and the hybrid scheme determine whether the membrane potential of a neuron crosses a threshold at the end of of the time interval between consecutive checkpoints. Threshold crossing can, however, occur within the interval even if this test is negative. Spikes can therefore be missed. The present work derives, implements, and benchmarks a method for perfect retrospective spike detection. This method can be applied to neuron models with affine or linear subthreshold dynamics. The idea behind the method is to propagate the threshold with a time-inverted dynamics, testing whether the threshold crosses the neuron state to be evolved, rather than vice versa. Algebraically this translates into a set of inequalities necessary and sufficient for threshold crossing. This test is slower than the imperfect one, but faster than an alternative perfect tests based on bisection or root-finding methods. Comparison confirms earlier results that the imperfect test rarely misses spikes (less than a fraction 1/108 of missed spikes) in biologically relevant settings. This study offers an alternative geometric point of view on neuronal dynamics. Keywords: State-space analysis, NEST, time-driven, event-driven, simulation, LIF neuron, differential geometry 1 INTRODUCTION In the last decade, considerable work has been devoted to improve the accuracy of simulators that are capable of efficiently simulating large networks of spiking neurons (Hansel et al., 1998; Mattia and Del Giudice, 2000; Shelley and Tao, 2001; Dehaene and Changeux, 2005; Morrison et al., 2007; Brette, 2007; D'Haene et al., 2009; van Elburg and van Ooyen, 2009; Zheng et al., 2009; Hanuschkin et al., 2010). The field is driven by the ideal of combining the capability to cope with the high-frequency of synaptic 1 J. Krishnan et al. Perfect spike detection via time reversal events arriving at a neuron in nature with a mathematically accurate implementation of the threshold process a wide class of neuron models is based on. Two classical schemes to simulate neuronal networks are the time-driven and the event-driven schemes (Fujimoto, 2000; Zeigler et al., 2000; Ferscha, 1996). Both schemes describe the state of the neurons by a set of variables and the action potentials as events that mediate the interaction between them. In a time-driven scheme, the state of a neuron is updated on a time grid defined by the simulation step (for a review see Morrison and Diesmann, 2008). After all neurons are updated, their membrane potential is checked for threshold crossings. If the membrane potential of a neuron is above the threshold at this checkpoint, a spike is delivered to all neurons it is connected to. Subsequently, a new iteration step begins. The step size stipulates how frequently occurrences of threshold crossings are inspected during the simulation. The choice of the step size a trade-off between spike accuracy and the speed of the simulation Morrison et al. (2007). Such grid-constrained simulations force each spike event to a position on the equidistant temporal grid spanned by the step size and therefore induce artificial synchronization of the network dynamics (Hansel et al., 1998; Shelley and Tao, 2001; Morrison et al., 2007; Brette, 2007; van Elburg and van Ooyen, 2009; Hanuschkin et al., 2010). In an event-driven scheme, the state of a neuron is updated only when it receives a spike. A central queue of events is maintained and each spike is inserted into this queue with its own time stamp. Upon update a neuron predicts when its next spike will occur in the absence of further input. This preliminary event is inserted into the queue and confirmed if it becomes due or removed when invalidated by further input. Efficient and elegant predicition methods have been developed for classes of neuron models without invertable dynamics (Brette, 2007; van Elburg and van Ooyen, 2009; D'Haene et al., 2009; D'Haene and Schrauwen, 2010; Ferscha, 1996). However, maintaining a central queue in a distributed simulation is challenging and may compromise the time performance of the simulator (Hanuschkin et al., 2010, for a detailed review see). A hybrid scheme circumvents the shortcomings of both these schemes by embedding a locally event- driven algorithm for each neuron into a globally time-driven scheme (Morrison et al., 2007). The arrival of a spike at a neuron introduces one additional update and check point. The dynamics of a given neuron is then propagated from incoming spike to incoming spike and eventually to the end point of the global timestep. If the membrane potential of a neuron is above the threshold at a local or global check point, the precise point of threshold crossing is determined in continuous time, and a spike is emitted. Next to their location on the time grid, in this scheme spike events carry a floating point offset. Thus, in contrast to an event-driven scheme, the hybrid scheme does not predict future spike times but identifies threshold crossings only retrospectively. Hanuschkin et al. (2010) demonstrate that the latter scheme is equally accurate as the former at lower computational costs. The hybrid scheme still has a loophole, however: spikes can be missed. The reason is that, as in the time-driven scheme, crossing of the threshold voltage is tested by inspecting whether the membrane potential of the neuron is above threshold at the end of a checkpoint (see Figure 1 for an illustration of the scenario). Nevertheless, the membrane potential V , evolved by the subthreshold dynamics, can be below threshold θ at two consecutive checkpoints t and t + h while a double, quadruple, etc. threshold-crossing occurred in between. The first crossing constitutes a missed spike. Symbolically, V (t + h) ⩾ θ =⇒ threshold crossing This is a provisional file, not the final typeset article (1) 2 J. Krishnan et al. Perfect spike detection via time reversal Figure 1. Illustration of undetected threshold crossing between two consecutive checkpoints t and t + h on the time grid. The short black vertical bar represents an incoming spike which causes an increase in the membrane voltage of the neuron, leading to a threshold crossing at tθ. The subthreshold dynamics, however, brings the voltage under threshold again at the next checkpoint. Since the test V (t + h) ⩾ θ yields false, the outgoing spike at tθ is missed. The red dots indicate points where the values of state variables are known. so the test V (t + h) ⩾ θ is a sufficient but not a necessary condition for the occurrence of a threshold crossing during the time interval ]t, t + h]. For future reference we call this test "standard test". The conceptual question therefore remains whether a globally time-driven scheme can formulated such that it detects every threshold crossing. Although Hanuschkin et al. (2010) argue that the loss of spikes of the standard test is not of practical relevance in natural parameter regimes, the availability of a method perfect by construction would free the researcher from inquietude and costly controls when faced with previously unexplored neuron models or network architectures. In this work we propose a new spike-detection method, which we term "lossless method" or "lossless test" because it is a necessary and sufficient condition for threshold crossing to occur in a given time interval. The method is based on state-space analysis and works with any neuron model with affine or linear subthreshold dynamics. It consists of a system of inequalities – some linear, some non-linear in the state-space variables – that together determine whether the initial state of a neuron will or will not reach threshold within the time interval until the next checkpoint. The lossless method replaces the standard test (1) in the time-driven and the hybrid scheme. Alone, it does not solve the problem of artificial synchronization the time-driven scheme suffers from. Hence the method is most meaningful within the hybrid scheme. Thanks to its perfect spike detection the lossless method can in fact be used to benchmark the hybrid scheme based on the standard test; Hanuschkin et al. (2010) use the method of D'Haene et al. (2009) for this purpose. In Section 2 we present the idea behind the lossless method for a general neuron model with an affine or linear dynamics, and develop its mathematical construction. Parts of this construction must be addressed on a case-by-case basis; therefore in Section 3 we provide a concrete implementation of the lossless method for the leaky integrate-and-fire model with exponential synaptic currents (Fourcaud and Brunel, 2002a), within the hybrid scheme. The method can be algorithmically expressed in different ways. We explore two alternative cascades of inequalities and assess their costs in terms of time-to-completion relative to each other and to the hybrid scheme based on the standard test (1). For the latter scheme we also assess the number of missed spikes in commonly considered network regimes. The hybrid scheme based on the lossless method delivers the desired exact implementation of the mathematical definition of the neuron model without any further approximation up to floating point precision. Frontiers 3 J. Krishnan et al. Perfect spike detection via time reversal Preliminary results have been published in abstract form (Kunkel et al., 2011; Krishnan et al., 2016). The technology described in the present article will be made available with one of next major releases of the open-source simulation software NEST. The conceptual and algorithmic work described here is a module in the long-term collaborative project to provide the technology for neural systems simulations (Gewaltig and Diesmann, 2007). 2 A TIME-REVERSED STATE-SPACE ANALYSIS 2.1 Idea: moving a surface backwards instead of a point forward Let us summarize the problem mentioned in the previous section. We assume that a neuron's state evolves according to three different dynamics: (a) an integrable subthreshold dynamic as long as the neuron's membrane potential is below threshold and there are no changes in input currents; (b) discrete jumps in the subthreshold dynamics at predetermined times, corresponding to incoming spikes or to sudden changes in external currents; these can be formally incorporated into the subthreshold dynamics (a) via delta functions; and (c) a "spike", i.e. an instantaneous jump of the membrane potential from threshold to a reset value, as soon as the potential reaches the threshold value. The jump may be followed by a refractory period in which the membrane potential remains constant at the reset value. Then the integrable dynamics takes place again. The advantage of the integrable dynamics is that the state of the neuron at a time t + h can be analytically determined by that at time t; here h can be negative or positive. The evolution can thus be calculated in discrete time steps, in particular in between times at which jumps (b) occur. The spike part of the dynamics, however, forces us to check whether the membrane potential V reached a threshold value θ within the timestep interval ]t, t + h]. We call this event threshold crossing (by "crossing" we also mean tangency). A sufficient condition for threshold crossing is that the membrane potential be above threshold at the end t + h of the time step: by continuity, it must have assumed the threshold value at some time in the interval ]t, t + h]. But this condition is not necessary: during the time step the potential may touch or surpass the threshold value and then go below it again, as in Figure 1, an even number of times. Its value is then below threshold at both ends of the time step, t and t + h. A test that only relies on the sufficient condition V (t + h) ⩾ θ – the standard test (1) – can therefore miss some spikes, leading to an incorrect dynamics. We need a test based on a necessary and sufficient condition. A necessary and sufficient condition for threshold crossing is that the trajectory of the state during ]t, t+h] intersect the hypersurface "membrane potential = threshold value". Translated into analytic geometry this means finding the solutions of a system of parametric equations – one representing the threshold hypersurface, the other the trajectory – and to test whether its solution set is empty (no threshold crossing) or not (threshold crossing). This idea is illustrated in Figure 2A for a two-dimensional state space. This system is usually transcendental and its solutions have to be found numerically. Unfortunately, numerical solutions typically rely on bisection algorithms (Press et al., 2007, ch. 9), involving an increasingly finer timestepping of the dynamics. This nullifies the advantage of having an integrable dynamics with coarse-grained time steps. The problem is that the test of intersection between trajectory and threshold tells us not only whether a threshold crossing occurs, but also the time at which it does. The test's high computational cost partly comes from delivering this additional information. This problem is avoided if we formulate a different geometric test that tells us whether a threshold crossing occurs but does not deliver the crossing time. This is a provisional file, not the final typeset article 4 J. Krishnan et al. Perfect spike detection via time reversal Mathematics often offer non-constructive proofs: "there exists a solution to the problem, but we do not know what the solution is". It turns out that this ignorance is exactly what we need in our problem. Instead of evolving the state of the neuron forwards in time, tracing a trajectory in state-space, and checking if and when it crosses the threshold hypersurface, we can evolve the threshold hypersurface backwards in time, sweeping a hypervolume, and check if and when it "crosses" the initial neuron state, which is a point. In other words we are testing whether a point belongs to a particular state-space region. The test for the intersection of a 1-dimensional curve with an (N − 1)-dimensional surface is replaced by the test for the membership of a point in an N-dimensional volume. This idea is illustrated in Figure 2B for a two-dimensional state-space. Mathematically the latter test translates into a system of inequalities that the initial state at t must satisfy if it does not cross threshold within ]t, t + h]. Even if transcendental functions appear in this system, we do not need to find their roots: we only need to test whether the inequalities are satisfied by simply inserting the value of the state-space variables in the functions. The equations corresponding to these inequalities represent the piecewise-differentiable boundary between the set of states that will cross the threshold within the timestep h – which we call spike region – and the set of those that won't – which we call no-spike region. Finding these inequalities in explicit form is the most important point of this method, and can be achieved with this heuristic procedure: I. Find the hypervolume swept by the threshold, in parametric form. This is done by representing the backpropagation of the threshold in state-space as a map between two manifolds: the product manifold threshold × time, and the state-space manifold. II. Find the boundary of the hypervolume, in parametric form. This is done by determining the placement of the images of the boundaries of the product manifold and, most important, of the images of the critical points of the map. The latter are the points at which the map becomes singular, thus mapping hypervolume elements into hypersurface elements (Spivak 1999, ch. 2; Choquet-Bruhat et al. 1996, § II.A.1). III. Transform the equations of the boundary above from parametric to implicit form. If this is not possible, the boundary can still be approximated by a triangularization, as finely as we please. IV. Finally, exclude the parts of the boundary that lie in the interior of the hypervolume (so technically not a boundary). Some guidelines to achieve this can be given based on the the convex properties of the hypervolume. In the following analysis we mathematically develop the first two steps in a general way for any neuron model with an affine dynamic. Let us point out that they can be generalized to other kinds of dynamics. The procedure for the last two steps depends on the particular neuron model, so we can only give general guidelines. Section 3 provides a concrete example of all steps. An advantage of this procedure is that it can be geometrically explained and needs to be carried out only once for any given neuron model. The advantage of this novel test is that its computational cost is distributed differently from the approach where the trajectory of the initial point is followed. In hand-waving terms, in the trajectory-based test the crossing time is immediately available from the parametrization of the orbit; the threshold-crossing condition is represented by computationally costly inequalities, because the crossing time, when it exists, can be read from them immediately. In the novel test the crossing time is a more complicated function of the hypervolume parametrization; the threshold-crossing condition in this case is represented by inequalities that are computationally less costly, because they do not explicitly deliver the crossing time. Knowledge of the latter requires an additional computational cost. The novel test thus allows us to avoid this extra cost when we check for the existence of a threshold-crossing; we only need to pay it if the crossing does occur. Frontiers 5 J. Krishnan et al. Perfect spike detection via time reversal Figure 2. Illustration of two exact methods to check whether an initial state (red dot) crosses the threshold (blue horizontal line) during the evolution from t to t + h. Upper panel: the idea of a root-finding method is to evolve the state forwards in time by a step ∆t = +h, and to check whether its trajectory (light-red curve) intersects the threshold. Such a method informs us of where the intersection occurs on the threshold, and at which time. Lower panel: the idea of the lossless method is to evolve the whole threshold "backwards in time" from t = h to t = 0 by a step ∆t = −h and to check whether its trajectory, which is a volume in state-space (light-blue region), contains the initial state, which is kept fixed. The threshold shifts and rotates as it evolves, and the trajectories of its individual points are unknown: this method does not inform us of when and where on the crossing it occurs, and is therefore computationally faster. From a mathematical point of view the two approaches are equivalent. The state-space technique is just a useful reformulation of the conventional trajectory view that is obtained by suitable mathematical manipulations. The conceptual device of propagating the threshold backwards in time is useful because it performs those manipulations in a transparent manner and gives them an intuitive dynamical meaning. This is a provisional file, not the final typeset article 6 θVt=0∆t=+ht=hIθVt=0∆t=−ht=h J. Krishnan et al. Perfect spike detection via time reversal 2.2 Mathematical preliminaries The final equations to be obtained, eq. (23), can be derived by concepts from vector analysis, Cartesian geometry, and functional analysis; but the derivation is lengthy. To shorten it we use concepts and terminology from affine spaces (Coxeter, 1969; Rockafellar, 1972; Nomizu and Sasaki, 1994; Artin, 1955; Porta Mana, 2011), and differential manifolds (Choquet-Bruhat et al., 1996; Burke, 1987, 1995; Simon et al., 1992; Nomizu and Sasaki, 1994; Marsden and Ratiu, 2007; Bossavit, 1991, 2002; Schouten, 1989; de Rham, 1984; Dodson and Poston, 1991; Ramanan, 2005). The state-space S of a neuron has a natural vector-space structure (an affine-space structure would also suffice), inherited from the physical quantities that define it: the membrane potential V and other N − 1 physical quantities I whose exact number and definition depend on the specific neuron model (e.g., I could represent currents or additional voltages, for example of different compartments). For our purpose it is useful to consider S as an N-dimensional differential manifold: its points {s} are the neuron states, and the quantities (I, V ) are coordinates V : S → R and I : S → RN−1 e.g., the membrane potential of a state s is V (s). These coordinates respect the vector structure of the state-space, i.e. V (s1 + s2) = V (s1) + V (s2) and likewise for I. ⊺ being Every hyperplane in the state-space is defined by an affine equation k the normal to the hyperplane, and −κ being the affine term. The inequality k s > −κ defines one of the two half-spaces delimited by the hyperplane. The threshold hyperplane is especially important: it is the set of states s whose membrane potential has the threshold value: s = −κ, the covector k ⊺ ⊺ Its equation k ⊺ s = κ in coordinates (I, V ) has coefficients V (s) = θ. ⊺ ⊺ = (0 s = κ on threshold, k , 1), ⊺ k κ = θ, s < κ below threshold. ⊺ k An affine transformation of the state-space onto itself, s (cid:55)→ Ms + m, where M is a linear transformation and m a state, maps each hypersurface and half-space k hypersurface and half-space k′⊺ s ⩾ κ′ with k′⊺ ⊺M−1, κ′ = κ + k ⊺M−1m = k ⊺ shows why it is a covector rather than a vector). (the transformation of the normal k (2) (3) (4) s ⩾ κ to a ⊺ (5) We now show that the integrable part of the neuron dyamics within a finite time step h is an affine transform. Similarly, the integrable part of of the neuron dynamics we consider an affine evolution determined by the equation (6) (7) 7 In a time interval h, this dynamics propagates an initial state s0 at time t into the final state s(t) = As(t) + q. s(t + h) = ehAs(t) +(cid:0)ehA − 1(cid:1)A−1q, s(t) = s0. Frontiers J. Krishnan et al. Perfect spike detection via time reversal (cid:18)B d (cid:19) This, for each h, is an affine transformation of the form (4). In coordinates (I, V ) the linear operator A and vector q have the block form A = (8) where B is an (N − 1, N − 1) matrix, α and β numbers, and the dimensionalities of the rectangular matrices ⊺, d follow accordingly. c 2.3 Derivation of the threshold-crossing condition q = α β c ⊺ , (cid:18)r (cid:19) , Let us mathematically summarize the first threshold-crossing test discussed in Section 2.1. We said that the state evolution (7) can be efficiently used in a time-step scheme in numerical simulations, but we need to test whether a threshold crossing occurred at some time t ∈ ]t, t + h]. A necessary and sufficient condition would be the existence of solutions of the trascendental equation in t t ∈ ]0, h], V(cid:2)etAs(t) +(cid:0)etA − 1(cid:1)A−1q(cid:3) = θ, (9) which corresponds to the intersection of the trajectory (7) and the threshold hyperplane (2); but it is a costly condition to test. We now develop the second kind of threshold-crossing test discussed in Section 2.1, according to the steps I–IV. Steps I and II are performed in full generality for an affine dynamics. Steps III and IV have to be solved on a case-by-case basis, so their analysis below is only a guideline; a concrete example on how to perform them is given in Section 3.3.3 and Section 3.3.4. 2.3.1 Hypervolume in parametric form Consider the threshold hyperplane V (s) = θ as an (N − 1)-dimensional manifold with coordinates x. It is embedded in state-space via the map x (cid:55)→ (I, V ) = (x, θ). (10) Each state (x, θ) on the threshold, when propagated backwards in time for an interval h, traces a curve in state-space (the yellow lines of Figure 3). The union of these curves is an N-dimensional product manifold, called the extrusion (Bossavit, 2003) of the threshold hyperplane. We can use coordinates (x, t) ∈ (RN−1 × [0, h]) on this manifold. Its mapping into state-space is given, with the help of eq. (7), by E : (x, t) (cid:55)→ e−tA (cid:19) (cid:18)x θ +(cid:0)e−tA − 1(cid:1)A−1 (cid:18)r (cid:19) β , (11) where t > 0 is the direction of the past. This map is analytic, but generally not an embedding because it can have self-intersections. We will see the significance of this in Section 2.4. It is not an immersion either because it can have singular points; these will be especially important for us because they constitute part of the boundary between spike and no-spike regions. See Figure 3 for a two-dimensional example. For fixed t, the map x (cid:55)→ E(x, t) is affine, and its image is a hyperplane representing the states on the threshold propagated backwards in time for an interval t. Combining it with eqs (3)–(5) we find that the This is a provisional file, not the final typeset article 8 V = θ, (cid:0)0 ⊺ 1(cid:1) ehA (cid:18)I (cid:19) V = θ +(cid:0)0 ⊺ TE(x, t) = (∂xE, ∂tE) = −e−tA 1(cid:1)(cid:0)1 − ehA(cid:1)A−1 (cid:18) I Bx + dθ + r ⊺ x + αθ + µ c ⊺ 0 (cid:18)r (cid:19) β . (cid:19) . (13) (14) (15) J. Krishnan et al. Perfect spike detection via time reversal backpropagated threshold has t-dependent normal and affine terms ⊺ t =(cid:0)0 κt = θ +(cid:0)0 1(cid:1) etA, 1(cid:1)(cid:0)1 − etA(cid:1)A−1 ⊺ ⊺ k (cid:19) . (12) (cid:18)r β ⊺ t s < κt determines the backpropagated half-space, which is below threshold. The inequality k 2.3.2 Hypervolume boundary in parametric form We must now find the boundary of the image of the map E. The latter is a closed set, being the image of a closed set under a continuous map; its boundary must therefore be the image of some points of the domain. Such points must either lie on the boundaries of the domain, RN−1 × {0} and RN−1 × {h}, or be critical points of E, or both, because E is differentiable. See the example of Figure 3. The images of the boundary are easily found from (11): one (image of t = 0) is the threshold hyperplane, ⊺ hs = κh, with coefficients given by (12). Explicitly, in coordinates the other (t = h) is the hyperplane k (I, V ), Let us find the image of the critical points of E. The tangent map of E at a point (x, t) is The tangent map is also denoted E∗ in some differential-geometry texts. Its determinant is the inverse ratio between a volume element at that point and its image in state-space (the sign determines their relative orientation). Hence this ratio vanishes at points where volume elements are mapped onto area elements (in other words, N linearly independent vectors in the domain are mapped onto N linearly dependent vectors), which is a feature of the boundary. See the two-dimensional example of Figure 3. Let us look for points where det TE(x, t) = 0. In eq. (15), the determinant of the exponential never vanishes, so we only have to consider the determinant of the matrix on the right. This is easily calculated by Laplace expansion along the last row, whose elements all vanish except the last. The cofactor of the last element is det I (modulo a sign). Hence det TE(x, t) = 0 ⇐⇒ c ⊺ x + αθ + β = 0 and the coordinates (x, t) of critical points satisfy ⊺ c x + αθ + β = 0, 0 ⩽ t ⩽ h. Frontiers (16) (17) 9 J. Krishnan et al. Perfect spike detection via time reversal This equation says that one of the coordinates x has an affine dependence on the remaining ones; let us call these y. For example, if cN−1 = 0, the equation above has the parametric solution x1 = y1, xN−2 = yN−2, xN−1 = −c1y1 + ··· + cN−2yN−2 + β + αθ . . . , cN−1 . (18) Denote this affine dependence by x(y). By taking the derivative of eq. (17) with respect to y we have ⊺ c ∂yx = 0 ⊺ , (19) a property we will use later. The locus of critical points in state-space is then given parametrically by a map Γ , found by substitution of eq. (18) in (11): Γ : (y, t) (cid:55)→ e−tA (cid:18)x(y) (cid:19) θ +(cid:0)e−tA − 1(cid:1)A−1 (cid:18)r (cid:19) β , 0 < t < h. (20) This locus has four important interrelated features: First: from the form of eqs (17) and (20), the locus of critical points is flat along N − 2 dimensions – corresponding to a fixed value of the coordinate t – and is curved normally to the direction t: it is an (N − 2)-ruled surface. Second: the points on the locus corresponding to fixed t belong to a hyperplane with coefficients (12), as is easily checked by substitution. In other words, the locus of critical points is the envelope of the ⊺ backpropagated threshold hyperplanes k t s = κt, eq. (12), at different times t, and its tangent hyperplanes have normals k ⊺ t given by (12). Third: considering the feature above for t = 0, the locus of critical points is tangent to the threshold hyperplane. Fourth: comparing the dynamics (6), the map for the threshold hyperplane (10), and the critical-point condition (17), we notice that the latter is also the condition for the trajectory (7) of a state s0 = (x, θ) on the threshold hyperplane to have an extremum in the membrane potential V . Hence, the locus of critical points is the trajectory of the intersection between the threshold and the V -nullcline. In view of the second property above, let us call the locus of critical points envelope, for brevity. 2.3.3 Hypervolume boundary in implicit form The next step is the elimination of the parameters (y, t) to express the envelope Γ as one or more implicit equations χ(s) = 0 defined on domains Dχ for the state-space coordinates s. This step can involve a transcendental equation, therefore we cannot give a general solution for it. In this regard, a criticism might be raised, which we immediately address: it looks like we have already taken great care in avoiding to solve eq. (9), only to face another difficult transcendental equation? The current problem is more manageable for three reasons. First, we observe that the equations needed to re-express the curved surface in explicit form are generally easier to handle than (9). Section 3.3.3 gives a concrete example. Second, the problem of finding the explicit form of the curved surface must be solved only once, whereas the solution of the original equation (9) has to be found for each initial state s0. Third, the equation for the curved surface This is a provisional file, not the final typeset article 10 J. Krishnan et al. Perfect spike detection via time reversal is easier to approximate numerically than the original one (9), again because the approximations do not depend on an initial state. Suppose we have found a function χ such that χ(s) = 0 is the envelope: {s χ(s) = 0, s ∈ Dχ} = {Γ (y, t) y ∈ RN−2, 0 ⩽ t ⩽ h}. (21) The condition above leaves χ completely undetermined (apart from smoothness requirements) outside of the envelope. Therefore, if possible, it is useful to extend the condition as follows. Since χ(s) = 0 is the ⊺ envelope of the backpropagated threshold hyperplanes k t s = κt, eq. (12), at different times t, it must ⊺ t s − κt < 0, for each t, correspond to the backpropagated be tangent to each of them. The inequalities k below-threshold half-plane, and it is useful to choose χ in such a way that the inequality χ(s) < 0, s ∈ Dχ, determines the side corresponding to the intersection of these half-spaces. An example is given ⊺ in Section 3.3.3. The differential dχ then is a positive multiple of k t are collinear because of the tangency condition). ⊺ t (note that dχ and k For the following, let us assume that χ has been chosen this way. 2.3.4 Boundary intersections and final system If a state does not cross the threshold hyperplane at any time t ∈ [0, h], then it must, by definition, belong for every t to the image of the half-space that lies below threshold, when this half-space is backpropagated by the time t. Mathematically, this is simply the statement of the equivalence ⊺ (Ms + m) < κ ⇔ (k ⊺M)s < (κ − k ⊺ k m) (22) established in Section 2.2, where M and m are the coefficients of the affine evolution by time t. By construction in the previous subsection, the inequality χ(I, V ) < 0 defines the region of intersection of all such backpropagated half-spaces, bounded by the envelope χ(I, V ) = 0. We just need to join to it ⊺ hs < κh, discussed the condition for the boundaries corresponding to the times t = 0 and t = h, V < θ, k in Section 2.3.1. The states that do not cross the threshold during the time interval [0, h] belong therefore to the no-spike region defined, in coordinates, by s0 = (I, V ) ∈ no-spike region ⇔ V < θ (cid:0)0 ⊺ 1(cid:1) ehA (cid:18)I (cid:19) V < θ +(cid:0)0 ⊺ 1(cid:1)(cid:0)1 − ehA(cid:1)A−1 (cid:18)r (cid:19) β and Frontiers (23a) (23b) 11 J. Krishnan et al. Perfect spike detection via time reversal and χ(I, V ) < 0, (I, V ) ∈ Dχ. (23c) Some inequalities in this system may turn out to be redundant, i.e. automatically satisfied if the remaining ones are, and can thus be dropped. Section 3.3.3 illustrates such a redundancy. 2.4 The region of missed spikes In the previous section we mentioned that the N-dimensional product manifold formed by the (N − 1)- dimensional threshold hyperplane and the time interval [0, h] presents self-intersections when mapped to the state-space. Two-dimensional examples of such a region are shown in Figure 3, lower panel, and in Figure 5, the region called S2. A state s in the self-intersection region corresponds to two or more different coordinates (x, t): s = E(x, t) = E(x′, t′), t = t′. (24) Note that x = x′, t = t′ is impossible for an affine transformation, since the first column of (15) can never vanish. The condition above simply means that s crosses the threshold at x after an interval t and at x′ after an interval t′. By continuity of the dynamics (6), one of the two must be a crossing from above, and one from below: this is exactly the scenario of double threshold crossing illustrated in Figure 1. Suppose we have a probability distribution for the initial states, for example one that is invariant under the dynamics (6). The probability of the self-intersection region P (X) is the probability that the initial state will lead to a double-crossing of the threshold, and therefore be missed. This fact will be used in Section 3.5.2 to estimate the number of spikes missed. 2.5 Convexity, approximations, optimization The affine dynamics (7) preserves affine combinations of solutions – and therefore convex combinations as well. If s1, s2 are two arbitrary initial states in the no-spike region, then their propagated states also satisfy V [s1(t)] < θ and V [s2(t)] < θ when 0 ⩽ t ⩽ h; and also the propagation of their convex combination λs1 + (1 − λ)s2, 0 < λ < 1 satisfies V [λs1(t) + (1 − λ)s2(t)] < θ, 0 < λ < 1, (25) i.e. it lies in the no-spike region. This proves that the no-spike region is convex. Convexity is important for approximations. If transforming the parametric equation for the envelope (20) into an implicit form (21) turns out to be analytically impossible, we can numerically find some points on the envelope and then approximate the latter by simplices constructed on these points, as finely as needed. In other words we can triangularize the envelope. Owing to convexity, the triangularization can be done completely on the side of the no-spike region (the corners of the simplices touch the envelope), or completely from that of the spike region (the barycentres of the simplices touch the envelope). This way we can formulate a test with no false positives, or one with no false negatives, or both. This is a provisional file, not the final typeset article 12 J. Krishnan et al. Perfect spike detection via time reversal The envelope is moreover a ruled surface, as shown in Section 2.3.2. The triangulation on the side of the no-spike region can therefore be conveniently chosen in such a way that one face of each simplex fully lies on the envelope. 3 IMPLEMENTATION EXAMPLE: LEAKY INTEGRATE-AND-FIRE NEURON WITH EXPONENTIALLY DECAYING POST-SYNAPTIC CURRENTS 3.1 The example model In the previous section we mathematically developed the idea of propagating the threshold backward in time in order to check whether a threshold-crossing occurs in a time-stepped dynamics. The derivation, valid for an affine subthreshold dynamics, is general and therefore also quite abstract; moreover, it involves a couple of mathematical steps (III and IV in Section 2.1) for which no general formulae can be given. To explain the idea in more concreteness and to give an example of how to face all its steps, we now apply the scheme to a simple but relevant model with a 2-dimensional state-space: the leaky integrate-and-fire neuron with exponentially decaying post-synaptic currents. This model has a homogeneous linear dynamics on a 3-dimensional state-space (Rotter and Diesmann, 1999), where the third coordinate is the input current. If this current is constant, the dynamics can be rewritten as a 2-dimensional affine one. Leaky integrate-and-fire models, despite their simplicity, approximate the behavior of real neurons with high accuracy (Rauch et al., 2003). The model with exponential synaptic currents captures important properties of real neurons: The postsynaptic potential has a finite rise and decay time and the membrane potential is a continuous function of time. Continuity avoids artificial synchronization, present in simpler models. Moreover, the model is to some extent analytically tractable. For short synaptic time constants, the mean firing rate (Fourcaud and Brunel, 2002a) as well as the linear response to small inputs (Schücker et al., 2015) can be obtained analytically. 3.2 Mathematical preliminaries: terms in block form The example model has a 2-dimensional state-space for a single neuron, defined by the post-synaptic current I and the membrane potential V , which are also our coordinates. Its subthreshold interspike dynamics (a) in Section 2 is affine: I(t), I(t) = − 1 τs I(t) − 1 τ V (t) = 1 C (cid:18)− 1 (cid:19) (cid:18) I V = V (t) + Ie, 1 C (cid:19) (cid:19)(cid:18) I V + (cid:19) , (cid:18) 0 1 C Ie 0 τs 1 C − 1 τ or in matrix form d dt (26a) (26b) where C is the membrane capacitance, τ is the membrane time constant, I is the synaptic input current and Ie is the external input current. The membrane potential V is subject to dissipation with time constant τ and integrates the post-synaptic current I. The latter decays exponentially with time constant τs. Typical values of the parameters are τ = 10 ms, C = 250 pF, τs = 2 ms, and the threshold θ = 20 mV. Frontiers 13 J. Krishnan et al. Perfect spike detection via time reversal Incoming spikes are incorporated in the equation for the current as Dirac deltas, and the external current Ie has jump discontinuities in time. In the timestepped evolution, such discontinuous events are implemented as instantaneous changes in the initial state s0 at each timestep. Hence we do not need to consider them explicitly in the equations above (Rotter and Diesmann, 1999). In terms of the block form of Section 2.2 we have The exponential of A is exp(−tA) = B = ( − 1 ⊺ = ( 1 C ), c τs ), d = ( 0 ), α = −1/τ, r = ( 0 ), Ie. 1 C  , β = 0 t τ e (27) (28)  (cid:0)e τs(cid:1)τ τs t e τs τ −e t C(τ−τs) t which determines the evolution of the neuron state s0 by an affine map as in eq. (7). Since the state space is 2-dimensional, the "hyperplanes" and "hypersurfaces" of Section 2 are straight lines and curves. In particular, the threshold hyperplane is a line; when propagated by a time t it maps onto a line with covector and affine term given by eq. (12), explicitly (cid:19) ⊺ t = k (cid:18)(cid:0)e− t τs(cid:1)τ τs κt = θ − Ieτ(cid:0)1 − e− t τ(cid:1) − t τ −e C(τ−τs) C e− t τ . , (29)  e− t τ V + Ieτ C 0 ⩽ t ⩽ h, (cid:18) For this model, testing for threshold-crossing by checking the intersection of a propagated state and the threshold line means solving the following transcendental system in t, given the initial state s0 = (I, V ): (1 − e− t τ ) + I τ τs C e− t τ − e− t τ − τs τs = θ, (30) for which we cannot find a solution in analytic form; we would have to resort to bisection algorithms. 3.3 The threshold-crossing condition 3.3.1 Hypervolume in parametric form The product manifold "threshold line × time interval" is in this case 2-dimensional, with coordinates (x, t). Here x is the value of the current at threshold crossing and t the corresponding time point. Its mapping E, eq. (11), to the state-space is E : (x, t) (cid:55)→ (I, V ) = e− t τs x, e− t τ θ + Ieτ C (1 − e− t τ ) + x τ τs C e− t τ − e− t τ − τs τs . (31) (cid:19) This map is shown in Figure 3: the x isocurves are yellow and in the lower panel they represent the trajectories of states terminating on the threshold; the t isolines are blue and represent the threshold line This is a provisional file, not the final typeset article 14 J. Krishnan et al. Perfect spike detection via time reversal propagated at different times. Each area element dx ∧ dt in the domain – the small rectangles in the upper plane – is mapped into an area element dI ∧ dV in the image. Note how these area elements are rotated and sheared. The thicker green curve is the set of singular points where the determinant of the tangent map vanishes: det(TE) = 0. Such points are singular because around them the images of the area elements get flattened to one dimension. In 3.4 we discuss the region of self-intersection bounded by the thick light cyan line, the thick dark blue line, and the thick green curved line. 3.3.2 Hypervolume boundary in parametric form The boundaries of the image of the map E must be, as explained in Section 2.3.3, a subset of the images of the boundaries of the domain, R × {0} and R × {h}, and of the envelope. The images of the boundaries, with general equations (13) and (14), in terms of (I, V ) follow The set of critical points of the map E is in this case a 1-dimensional curve, given in parametric form by V = θ, V = e h τ θ + Ieτ C (cid:18) (cid:16) θC τ Γ : t (cid:55)→ t τs e − Ie , τ(cid:1) + h Iτ C (cid:0)1 − e (cid:17) τ e τs e τse− h (cid:16)θC h τs h τ − e τ − τs (cid:17) t τs t τ − τse τ − τs τ C − Ie τ + τ C Ie ; (cid:19) . (32) (33) the coordinate y of the general form (20) do not exist in this case, because the threshold is 1-dimensional. Figure 3 visualizes the set by the green curve. 3.3.3 Hypervolume boundary in implicit form The next step in our procedure is to convert the parametric equation (34) of the envelope into an explicit or implicit equation for the coordinates (I, V ). As Section 2.3.3 does not provide a general algorithm, below we illustrate the process using our example model. By equating the first component of eq. (34) to the coordinate I and solving for t, we find t = −τs ln − Ie (cid:104)(cid:16) θC (cid:16)θC τ , /I (cid:105) (cid:17) (cid:17) ⩽ 1, subject to the condition for I e− h required for 0 ⩽ t ⩽ h and a real logarithm. Substituting eq. (35a) into the V coordinate of eq. (34) we find τs ⩽ 1 I − Ie τ τ(cid:2)(cid:0) θC τ − Ie (cid:1)/I(cid:3)1− τs τ − τs τ − τs V = τ Ie C + τ I C Frontiers (34) (35a) (35b) (36) 15 J. Krishnan et al. Perfect spike detection via time reversal Figure 3. Example of map E : (x, t) (cid:55)→ (I, V ), eq. (11), in two dimensions. Upper panel: the abstract manifold with coordinates (x, t) corresponding to the values of current and time at threshold crossing. Lower panel: the image of the map in state space. Thick green curve corresponds to the set of singular points where det(T E) = 0, eq. (9). Yellow lines, constant x, are trajectories of states ending on the threshold. Violet lines, constant t, are snapshots of the threshold moving "backwards in time". The thicker violet and blue lines correspond to the boundaries t = 0 and t = −h. For t = 0 we have I = x, V = θ. subject to the condition (35b). Let us analyse this equation, in view of its extension to an inequality of the form χ(I, V ) < 0 as required in Section 2.3.3. First, we observe that τ e t τs t τ − τse τ − τs ⩽ 1 for 0 ⩽ t ⩽ h. This is a provisional file, not the final typeset article (37) 16 x−h0tIθV J. Krishnan et al. Perfect spike detection via time reversal The inequality can be proven by studying the derivative of the fraction with respect to t. The derivative is always negative in the range above and the only maximum of the fraction is the value unity assumed at t = 0. Inspection of equation (36) and of its parametric form (34) shows that we must consider three cases: Ie ⋚ θC/τ, i.e. whether the external current is smaller or larger than the rheobase current this is the current necessary to reach threshold in an infinite time starting from any state with I ⩽ 0. • If Ie < Iθ, then I is restricted to Iθ = θC/τ ; 0 < Iθ − Ie ⩽ I ⩽ e h τs (Iθ − Ie). (38) (39) In this case, using the inequality (37), the V component of the envelope (34) is always smaller than the threshold: τ(cid:2)(Iθ − Ie)/I(cid:3)1− τs τ − τs τ − τs • If Ie > Iθ, then I must be negative and restricted to + V ≡ τ Ie C τ I C ⩽ θ when Ie < Iθ. h τs (Iθ − Ie) ⩽ I ⩽ Iθ − Ie < 0. e (40) (41) τ(cid:2)(Iθ − Ie)/I(cid:3)1− τs τ − τs In this case, using the inequality (37), the V component of the envelope (34) is always larger than the threshold: V ≡ τ Ie C τ I C + (42) • If Ie = Iθ, the envelope degenerates to a point: Γ (t) = (0, θ), which is the limit point reached in infinite time from any initial point in the no-spike region; this is the geometric interpretation of the equality of external and rheobase currents. τ − τs ⩾ θ when Ie > Iθ. The function representing the envelope, eq. (21) of Section 2.3.3, has in this case the explicit form χ(I, V ) := V − τ Ie C − τ I C τ(cid:2)(Iθ − Ie)/I(cid:3)1− τs τ − τs τ − τs , with (I, V ) ∈ Dχ := (cid:104) Iθ − Ie, e (cid:105) × R. h τs (Iθ − Ie) (43) If we calculate its differential, as discussed in Section 2.3.3, we find that the latter is a positive multiple of the differential of the backpropagated threshold for each t. This is true in each of the three cases above. Consequently, the inequality χ(I, V ) < 0, (I, V ) ∈ Dχ always includes the no-spike region. In summary, we arrive at an analytic, implicit equation for the curved boundary (43). The expression is a transcendental function in I, owing to the generally irrational exponent 1 − τs/τ, but it is used in an inequality, hence we do not need to find its roots. This is in contrast to the original transcendental equation for the threshold-crossing condition (30), which requires a bisection algorithm to find its solution. Frontiers 17 J. Krishnan et al. Perfect spike detection via time reversal 3.3.4 Boundary intersections and final system We can now assemble the system of inequalities defining the no-spike region consisting of the boundaries (32) and (33), and the envelope (43). With some rearrangements, simplifications, and the introduction of two new functions f, b, the condition reads: s0 = (I, V ) ∈ no-spike region ⇔ V < θ and and V < fh,Ie(I) := e h τ θ + τ Ie C h (cid:0)1 − e τ(cid:1) + τse− h τ(cid:2)(Iθ − Ie)/I(cid:3)1− τs τs h τs h τ − e τ − τs e τ I C V < bIe(I) := τ Ie C + τ I C τ − τs τ − τs if I ∈(cid:104) Iθ − Ie, e (cid:105) h τs (Iθ − Ie) (44a) (44b) (44c) Figure 4 illustrates the system for the two cases Ie < Iθ and Ie > Iθ. In the former case all three inequalities are necessary; in the latter, as well as for Ie = Iθ, the last inequality is automatically enforced by the first because its right-hand side is larger than the threshold θ (see eq. (42)). 3.4 The region of missed spikes In Figure 3, lower panel, the trajectories of several states during a timestep h and ending on the threshold line are represented by yellow curves. In that figure we can identify a region were such trajectories self- intersect: it is bounded by a segment of the thick light blue line, a segment of the thick dark blue line, and a portion of the thick green curved line. Trajectories with initial states in this region must therefore cross the threshold twice during the interval ]t, t + h]. As explained in Section 2.4, their threshold-crossing is not detected by the sole condition V [s(t + h)] ⩾ θ. All states in this region thus generate spikes that are missed by the standard test (1). This region is crucial for the comparison of the performances of schemes implementing the present lossless method and schemes relying on the standard test (1). This comparison is quantitatively made in Section 3.5.3. This is a provisional file, not the final typeset article 18 J. Krishnan et al. Perfect spike detection via time reversal Figure 4. System of inequalities, eq. (44), determining the no-spike region. The colored areas represent the complementary inequalities of the system (44), so the solution of that system is the white area. The red region delimited by the horizontal line corresponds to the first equation of the system, the blue region delimited by the inclined line to the second, and the yellow region delimited by the curved line to the third. Upper panel: case Ie < Iθ, all inequalities necessary. Lower panel: case Ie > Iθ, the third inequality is redundant. Frontiers 19 0Iθ−Ieeh/τs(Iθ−Ie)θVIe<IθIθ−Ieeh/τs(Iθ−Ie)0IθVIe>Iθ J. Krishnan et al. Perfect spike detection via time reversal 3.5 Optimization, time performance, and accuracy 3.5.1 Numerical implementation and optimization In the last section we arrived at the system of three inequalities (44) that determines whether or not the current state (V, I) will cross the threshold within the timestep h. The state will cross the threshold if the system is not satisfied, and it will not cross the threshold if the system is satisfied. This system of inequalities constitutes the lossless method in the present model. The system requires the current timestep h, state (V, I), and external electric current Ie as inputs (cf. Section 2). The timestep h is the minimum between the global timestep and the time interval up to the next input coming from other neurons, hence it can differ every time the test is called. The external electric current Ie may also vary, stepwise, during the simulation, hence it may also be distinct at each call of the test. If the system of inequalities is satisfied, thus predicting the absence of a spike within a timestep h, then the state of the neuron is evolved by applying the propagator (7) with (27)–(28), leading to a new state, and the procedure starts again. If the system is not satisfied, thus predicting the occurrence of a spike within h, it is then necessary to compute the time tθ at which the threshold is crossed. This calculation, explained in Appendix A, is made by interpolation between the current state and time, and the state and time tmax at which the membrane potential would reach its maximum if allowed to increase above threshold. Once tθ is calculated, a spike is emitted, communicated to the postsynaptic neurons, and the membrane potential is reset to Vreset for a refractory period. When this refractory period is over the procedure starts again. In the evolution loop just described, the membrane potential is reset to a value below threshold as soon as it crosses the latter. Thus no initial state can have V ⩾ θ. This means that the first inequality (44a) in the system is always satisfied and can be dropped. Only inequalities (44b) and (44c) have to be assessed, leading to the reduced system (45) (cid:40) V < fh,Ie(I), V < bIe(I). We first discuss the case Ie < Iθ ≡ θC/τ. The geometric meaning of the inequalities above is illustrated in Figure 5. The figure shows four regions: NS1, NS2, S1, S2. The no-spike region is the union of NS1 and NS2, the spike region the union of S1 and S2. In the figure they are separated by a thick line, partly curved and black, partly straight and blue. Subregion S2 is particularly important: it is the region of missed spikes discussed in Section 3.4, corresponding to the self-intersection region of Figure 3, lower panel. It contains those states that lead to spikes missed by schemes that rely on the standard test (1). Region S1 is separated from S2 by a dashed blue line, and from NS1 by a continuous blue line, the continuation of the dashed one. This partly dashed, partly continuous blue line corresponds to the equation V = fh,Ie(I). Hence if the inequality V < fh,Ie(I) is not satisfied then the initial state is in region S2 or on its blue boundary, and there will be a spike. If the inequality is satisfied the state could be in S2 – spike – or NS1 ∪ NS2 – no spike; an undetermined case. This inequality requires modest computational costs because it is linear in I and Ie and involves exponentials of h. Regions S2 and NS2 are separated by a black curve: . Hence if the inequality V < bIe(I) is not satisfied the initial state is in S2 or on its boundary, and there will be a spike. If the inequality is satisfied the initial state is either in NS2, or in NS1 with Iθ − Ie < I < e τs (Iθ − Ie), and no spike will occur. this is the envelope, corresponding to the equation V = bIe(I) for I ∈(cid:104) Iθ − Ie, e h τs (Iθ − Ie) (cid:105) h This is a provisional file, not the final typeset article 20 J. Krishnan et al. Perfect spike detection via time reversal Figure 5. State-space subregions formed by the intersections of the reduced inequalities (45): V < fh,Ie(I) (straight blue line, partly dashed partly continuous) and V < bh(I) (black curve), and by the auxiliary inequality (47): V < gh,Ie(I) (dot-dashed straight purple line). The spike region is S1 ∪ S2, the no-spike region is NS1 ∪ NS2. The subregion S2 contains all states that emit spikes undetected by the standard test (1); they are detected by the lossless method. The computationally most expensive inequality is V < bIe(I) because it involves irrational powers of I and Ie. It is advisable to avoid its direct computation as often as possible by pre-testing a linear inequality. In Section 2.5 we discussed how such a pre-test is indeed possible thanks to the convexity of the no-spike region. There, we argued that the curved envelope can be approximated by triangular hypersurfaces, which simply reduce to one straight segment in the present two-dimensional case: this is the dot-dashed red line in Figure 5, separating NS1 and NS2. This line has equation V = gh,Ie(I) with gh,Ie(I) := θ + h τs τse τ − τs τ C I + e h τ (Iθ − Ie), τ C and the corresponding inequality has the same computational costs as V < fh,Ie(I). V < gh,Ie(I) (46) (47) If the auxiliary inequality V < gh,Ie(I) is satisfied, the initial state is in NS1 and V < bIe(I) is also satisfied. It is therefore convenient to test the auxiliary inequality before the computationally costly one, which can be discarded if the test is positive. Figure 5 suggests that this test might be positive for the Frontiers 21 Iθ−Ieeh/τs(Iθ−Ie)IθVNS1S1NS2S2fh,Ie(I)gh,Ie(I)b(I) J. Krishnan et al. Perfect spike detection via time reversal majority of initial states because region NS1 is much wider than NS2. This possibility would be very advantageous, but we now argue that it should be verified by a dynamical analysis. The average time cost of the algorithm in a long simulation is given by(cid:80) The system (45) can be translated into a computational algorithm in several different ways, depending on the order of evaluation of its two inequalities and of the auxiliary inequality (47). In simplified terms, such an algorithm consists in a sequence of tests – variously implemented as if, and, or constructs – for finding the initial state in space-time regions R1, R2, and so on. The order of these tests is important. i pici, where pi is the frequency with which states are found in region Ri, which we call "occupation frequency", and ci is the cumulative time cost of the test for region Ri. This time cost ci is cumulative in the sense that all tests up to the (i − 1)th must have been performed, with false outcomes, to arrive at the test for Ri. The efficiency of an algorithm therefore depends on the mathematical form of the inequalities defining a region and on the occupation frequencies of the regions, determined by the dynamics. These two factors can be extrapolated by a theoretical analysis, or more practically measured by running long test simulations with typical network setups corresponding to the cases one is interested in. We now try to determine the most efficient algorithm for the present case. Region S1 is the least costly, because bounded by one line and therefore involving one inequality linear in I; then region NS1, bounded by two lines involving two linear inequalities; and finally regions NS2 and S2, bounded by the curve that involves rational exponentiation. For this example model, we tried different orderings but show here only two possible extreme cases to illustrate that there is no significant difference in the computational cost. Algorithm 1 Algorithm 2 bool is_spike(h): pre-compute gh,Ie(I) if V ⩽ fh,Ie(I) and V < gh,Ie(I) then else if V ⩾ gh,Ie(I) then else if V ⩾ bIe(I) then return false return true return true else return false bool is_spike(h): if V ⩾ gh,Ie(I) or [V ⩾ fh,Ie(I) and V ⩾ bIe(I)] then return true else return false Algorithm 1 is based on the assumption that the occupation frequency of a subregion is proportional to that subregion's relative size. If we check the two largest first, in the order NS1, S1, S2, NS2, we are therefore more likely to exit the test in its first if branches. The value of gh,Ie is used in two if branches, so it is computed just once and saved before the if sequence in order to save some computations. Algorithm 2 uses a composite or-and condition rather than several ifs. Assuming left-to-right evalu- ation, the algorithm corresponds to testing first S1 (left side of or), then S2 (right side of or), either case leading to a spike. If neither is true, no further tests are necessary because the state must necessarily be in the no-spike region. The test for S2 is made less costly on average by using the auxiliary inequality (47). This algorithm uses one test less overall than the previous one, but it may require one more test on average, if NS1 is the region with highest occupation frequency. The left-to-right evaluation assumption does not always hold in modern processors, which build their own statistics to optimize the test order of logical constructs. This is a provisional file, not the final typeset article 22 J. Krishnan et al. Perfect spike detection via time reversal Figure 6. (A) Schematic of the simulation setup used to calculate occupation frequencies and to compare the hybrid scheme with lossless method Algorithm 1, with lossless method 2, and with the standard test. The neuron model (empty red circle), implementing one of the three schemes, receives input from an external current and from one excitatory (E) and one inhibitory (I) Poisson generator. The total input has mean µ and variance σ2. (B) Sample of membrane-potential dynamics for µ = 15 mV and σ2 = 25 mV2. The analysis assumed Ie < Iθ ≡ θC/τ. The reduced system (45) and the Algorithm 1 and 2 are, however, also valid in the case that Ie ⩾ Iθ, corresponding to the lower panel of Figure 4. In this case subregions NS2 and S2 do not exist below the threshold, and the inequalities V < bIe(I) and V < gh,Ie(I) are always satisfied when V < θ; they always evaluate to true in both algorithms. Both algorithms therefore correctly distinguish spiking from non-spiking states in this case, although they become inefficient owing to the additional superfluous evaluations of bIe(I) and gh,Ie(I) for spiking states. We decide not to modify them in the present work because the case Ie ⩾ Iθ is unusual in real applications. More efficient algorithms for this case can be designed by interested readers following the guidelines just given in this section. 3.5.2 Occupation frequencies We want to measure the occupation frequencies in the typical case of a neuron embedded in a recurrent network, receiving fluctuating synaptic input. The setup for this simulation is illustrated in Figure 6A and its formulae explained in Appendix B. One neuron is coupled with strengths J and −J to an excitatory and an inhibitory Poisson generator and also receives a constant external current. The Poisson generators each mimic an excitatory and an inhibitory population. The neuron thus receives a fluctuating input current having average µ and variance σ2. In this instance the code of the simulation includes a subroutine that informs us of the current state every time the lossless method test is called, without altering the test or the dynamics. Figure 7A gives a visual idea of the occupation frequencies for µ = 15 mV, σ2 = 25 mV2, and J = 0.1 mV (all expressed in volts through multiplication by a resistance of τ /C = 40 MΩ), which correspond to the case Ie < Iθ. These values correspond to a composite average input of 250 000 spikes/s, and a total average input current of 400 pA. This presynaptic input makes the neuron fire at an average rate of 7 spikes/s. A sample of its membrane dynamics is shown in Figure 6B. Subregion NS1 has the overwhelmingly largest occupation frequency. The other three subregions have actually very small Frontiers 23 IAF neuron J. Krishnan et al. Perfect spike detection via time reversal Figure 7. (A) Frequency density of states over state space at each call of the threshold-crossing test, for network parameters µ = 15 mV, σ2 = 25 mV2, J = 0.1 mV. The colourbar is in units of 6 × 109 mV−1 pA−1 (obtained from total number of events × area element). The dotted purple line is the threshold θ. The only visible region in this plot is NS1. and the horizontal blue line is the boundary between no-spike and spike regions. The spike region S1∪ S2 and subregion NS2 are not visible on this scale because of the exceedingly small average timestep h = 4 × 10−3 ms. To discern them we need to zoom in, as done in upper panel (B): the curved envelope and the two straight lines that separate S1, S2, and NS2 extend horizontally and vertically for just about 1 pA and 10−5 mV. (B) lower panel. In contrast, for a much larger timestep h = 5 ms the three boundaries would have a larger extension, about 5 500 pA and 20 mV, and be discernible in plot (A). areas, as clear from the axis ranges of Figure 7B, owing to the very small value of the average timestep, h = 4 × 10−3 ms, given by the inverse of the input rate in this hybrid scheme. A more precise comparison of the occupation frequencies of the four regions N S1, N S2, S1, S2 is shown in Figure 8 for several combinations of three network parameters, producing different dynamic regimes. The parameters are the average µ, the variance σ2 of the input current, and the presynaptic coupling strength J (all expressed in volts through multiplication by a resistance of τ /C = 40 MΩ). The values of the parameters (µ, σ2, J) include typical realistic cases as well as some extreme cases, like unusually high coupling strengths. Each panel of Figure 8 shows the occupation frequencies for a set of dynamical regimes with constant J, β and several σ2. The panels in the last row correspond to the case Ie ≡ µC/τ ⩾ Iθ ≡ θC/τ, or µ ⩾ θ, in which subregions NS2 and S2 do not exist below threshold. It is important to remember that the boundary and size of the regions of Figure 5 vary with the timestep h, which is a parameter of the simulation scheme, not of the dynamics per se. In an event-driven or hybrid scheme, this step varies inversely with the event input rate, which for Poisson input generators is proportional to σ2/J 2. As a consequence, the frequencies displayed in Figure 8 are not determined by the Liouville distribution of the dynamics (26) alone, but also by the details of the numerical-implementation scheme. The dependence of the boundaries on h is illustrated in Figure 7B. As h decreases, the line V = fh,Ie(I) and the auxiliary line V = gh,Ie(I) gets closer to the threshold, and the subregions S1, S2, NS2 disappear. This is plausible since in the limit of h = 0 we are not evolving the initial state at all. As h increases the point of tangency between the envelope V = bIe(I) and the line V = fh,Ie(I) moves to This is a provisional file, not the final typeset article 24 p J. Krishnan et al. Perfect spike detection via time reversal increasingly lower voltages and higher currents; subregion S2 takes over S1 and subregion NS1 becomes wider. For typical timestep values of several milliseconds, though, subregions S1, S2, NS2 are still very small. A rough estimate of the dependence of the areas of the bounded regions S2 and NS2 on the parameters (µ, σ2, J), for µ < θ, can be obtained by looking at Figure 8 and considering that these areas together form a triangle with vertices (Iθ − Ie, θ), with If such that fh,Ie(If ) = θ. (48) This triangle has base If − (Iθ − Ie) and height θ − bIe[eh/τs (Iθ − Ie)]. Expressing h and Ie in terms of µ and σ2 using eqs (53), where h is inversely proportional to the input rate rI + rE, we find (cid:0)eh/τs (Iθ − Ie), bIe[eh/τs (Iθ − Ie)](cid:1), (If , θ), (cid:20) 1− τ eh/τ − τseh/τs (cid:21)(cid:20)eh/τs (τ − τs) (1 − eh/τ ) (cid:21) (θ−µ)2 areas of S2 and NS2 ∝ C τ τ J 2 σ2 . (49) When τs ≲ τ and J 2 ≲ σ2 a Taylor expansion in J 2/σ2 to fourth order gives a good approximation, with a relative error below 10 %: τs (eh/τ − eh/τs) τ − τs with h = −1 (cid:34) (cid:18) J 2 (cid:19)3 σ2 τ 2 4τ 2 s (cid:18) J 2 (cid:19)4 τ 2 (τ + τs) + 8τs3 σ2 (cid:19)5(cid:35) (cid:18) J 2 σ2 + O . (50) areas of S2 and NS2 ∝ C τ (θ − µ)2 This approximate formula shows that subregions NS2 and S2 grow with the square of the input mean µ and with the third or fourth power of the ratio J 2/σ2. Recall that these regions do not exist for µ ⩾ θ. The occupation frequencies do not depend on the areas alone, however, but also on the dynamics, as explained in the previous section. We can identify several other dynamical mechanisms for their dependence on the parameters (µ, σ2, J): • an increase in mean input µ leads to more frequent threshold crossings, thus frequently bringing the voltage to its reset value, underneath subregions NS2 and S2. The occupation frequencies of these subregions may therefore decrease with µ even though their areas grow with µ; • for low mean input µ, an increase in variance σ2 means a higher chance of high-V regions, and thus an increase in the occupation frequencies of NS2 and S2, even though their areas shrink with σ2; • for mean input µ close to the threshold, an increase in the variance σ2 leads to more frequent threshold crossings, and may thus increase occupation frequency of S2 with σ2, even though its area shrinks with σ2. The occupation frequency of subregion NS1 (green circles) dominates all others, varying from 90 % to 100 % depending on the network parameters. Subregion S1 (yellow crosses) follows in order of frequency and is the most frequently visited between the two spike subregions. Subregions NS2 (blue squares) and S2 (red stars) are scarcely visited for lower synaptic amplitudes, with frequencies from 0 to 10−6; and slightly more often at higher synaptic amplitudes (frequencies from 0 to 10−2). 3.5.3 Time performance and accuracy of a hybrid scheme based on the lossless method The average time costs of Algorithm 1 and Algorithm 2 within a hybrid scheme can be assessed by real-time simulation measurements. The average time cost of a hybrid scheme based on the standard test (1) can also be assessed in the same way for comparison. We therefore compare the two algorithms of the lossless method and the sufficiency test in this section. Frontiers 25 J. Krishnan et al. Perfect spike detection via time reversal Figure 8. Occupation frequencies of the four subregions NS1 (green circles), S1 (yellow crosses), NS2 (blue squares), S2 (red stars) of Figure 5, for various sets (µ, σ2, J) of input-current mean and variance, and synaptic strength. Each panel shows the frequencies vs current variance for fixed current mean and synaptic strength. The columns have the same J, ranging from 0.1 mV (leftmost) to 5 mV (rightmost). The rows have the same µ, ranging from 10 mV (top) to 22 mV (bottom). The frequencies were measured from N samples, depending on (µ, σ2, J). The dotted lines in each plot show the inverse number of samples 1/N for that network regime. The thickness of the segments connecting the data points equals one standard deviation. The shaded regions show where the limiting frequencies (N → ∞) are expected to lie with 87 % probability, using a Johnson-Dirichlet model with parameter k = 0.05 determined by posterior maximization (Johnson, 1932; Zabell, 1982; Good, 1966; Bernardo and Smith, 2000, § 3.2.5). The frequencies of region S2 (red stars) are particularly important: they are the frequencies of spike-misses of the standard test (1). This is a provisional file, not the final typeset article 26 10-810-710-610-510-410-310-210-1100frequencyJ=0.1mVJ=0.5mVNS1S1NS2S2J=1.0mVJ=5.0mVµ=10mV10-810-710-610-510-410-310-210-1100frequencyµ=14mV10-810-710-610-510-410-310-210-1100frequencyµ=18mV101418222630σ2/mV210-810-710-610-510-410-310-210-1100frequency101418222630σ2/mV2101418222630σ2/mV2101418222630σ2/mV2µ=22mV J. Krishnan et al. Perfect spike detection via time reversal The basic setup is the same as for the frequency analysis of 3.5.1, explained in Appendix B, with network parameters (µ, σ2, J). The only difference is that in the present case the code does not include the subroutine that informs us of the frequencies, which would otherwise increase and bias the real-time durations of the simulations. Three instances of the basic setup are prepared: in the first the neuron is modelled by a hybrid scheme with lossless method Algorithm 1, in the second the neuron is modelled by a hybrid scheme with lossless method Algorithm 2, and in the third the neuron is modelled by a hybrid scheme with the standard threshold-crossing test (1). The various random-number-generator seeds of the three instances are exactly the same, so that the three neurons receive exactly the same input, spike-for-spike; this is essential for a fair comparison between the three schemes. The three instances are run for a long time (2 × 106 ms) and repeated (in parallel) for several times (10), enough to collect reliable statistics. The statistics are collected for the same sets of (µ, σ2, J) values as in the frequency analysis. The total real-time length of a simulation depends, in all three schemes, on how often the checkpoints and threshold-crossing tests occurs, and this in turn depends on the presynaptic input frequency, as already discussed. Rather than showing the results for all sets of parameters (µ, σ2, J), which in this case are not very informative, we show in Figure 9 those with the (J, µ) values that yield the slowest and fastest performances. The average computation costs of Algorithm 1 and Algorithm 2, which embody the lossless method, turn out to be very similar – within each other's standard deviations – and basically identical in comparison with the cost of the scheme based on the standard test (1). Both are slower than the hybrid scheme with the standard test: from around 33 % slower in the case of high-activity regime with frequent incoming spikes (µ = 18 mV, J = 0.1 mV), to around 8 % slower in the case of low activity regime and infrequent incoming spikes (µ = 10 mV, J = 5 mV). These are the extremes shown in Figure 9. In most other sets of network parameters the hybrid scheme with lossless method was around 20 %–25 % slower than the standard hybrid scheme with threshold-crossing test (1). As mentioned in 3.4, subregion S2 contains all states for which the standard threshold-crossing test misses a spike. The occupation frequencies of this subregion are therefore a direct measure of the number of spike missed, per neuron, by the standard hybrid scheme. They are shown as red stars in Figure 8 for the various sets of network parameters (µ, σ2, J). The frequency of missed spikes does not have a simple monotonic dependence on the three parameters, owing to the interaction of several mechanisms, discussed in Section 3.5.2. An increase in synaptic coupling J or input current µ generally leads to more frequently missed spikes because the neuron spikes more often overall. For very high – suprathreshold – input currents, however, the frequency decreases again until no spikes are missed anymore; this change in trend happens because the checkpoints become more frequent. Regimes of low J are diffusion-like processes, where frequent arrival of synaptic events does not lead to missed spikes. Regimes of high J are shot-noise processes, where sudden and infrequent arrival of synaptic events leads to suprathreshold excursions and missed spikes. Separate simulations show that the hybrid scheme based on the lossless method, with either algorithm, reproduce the analytic solution of the neuron model within floating point precision. For biologically realistic synaptic couplings, J < 1 mV, the standard hybrid scheme is 33 % faster than the scheme with the lossless method, and misses less than 1 spike every 106 test calls, per neuron; for very low couplings J < 1 mV this figure even becomes less than 1 spike every 108 test calls. For synaptic couplings J > 1 mV the standard hybrid scheme starts to miss more spikes, reaching even 1 missed spike every 500 test calls for subthreshold average input currents; and it is only 8 % faster than the hybrid scheme with the lossless method. From these figures a user can decide to use the hybrid scheme with the standard or the lossless test, depending on the desired balance of accuracy and speed. Frontiers 27 J. Krishnan et al. Perfect spike detection via time reversal Figure 9. Computational costs of the hybrid scheme with lossless method, Algorithm 1 (red), lossless method, Algorithm 2 (green), and standard threshold-crossing test (1) (blue). (A) µ = 18 mV, J = 0.1 mV describes regimes where the lossless method has the highest increase in computational cost, around 33 %. These are regimes of high activity and frequently incoming spikes. (B) µ = 10 mV, J = 5 mV describe regimes where the lossless method has the lowest increase in computational cost, around 8 %. These are regimes of low activity and infrequently incoming spikes. The data come from 10 simulations of 2× 106 ms simulation-time each. This is a provisional file, not the final typeset article 28 (mV) J. Krishnan et al. Perfect spike detection via time reversal 4 SUMMARY AND DISCUSSION We here present a general method to solve the threshold-crossing detection problem for an integrable, affine or linear neuronal dynamics. The method is based on the geometric idea of propagating the threshold plane backwards in time and to determine whether the swept volume contains the initial state, rather than propagating the initial state forward in time to check whether it crosses the threshold. These two procedures are obviously mathematically equivalent, but they distribute the computational load in different ways. The forward-propagation of the state looks for the value of the crossing time; if no such value exists, it means there is no threshold crossing. The backward-propagation of the threshold first tests whether a crossing time exists at all, without yielding its value; the latter is calculated afterwards. The different distribution of computational load in the two procedures can be explained geometrically and algebraically. The first procedure geometrically checks for the intersection of a curve (the state trajectory) with a hypersurface (the threshold). Algebraically, this corresponds to finding the roots of a system of equations, often transcendental. The second procedure geometrically checks for the "intersection" of a point (the state) with a hypervolume (the threshold trajectory), i.e. the inclusion of the former in the latter. Algebraically, this corresponds to testing a set of inequalities. The latter procedure is more efficient for numerical computations, because it only relies on inequality tests – in which the presence of transcendental functions is much less costly than in an equation that needs to be solved for a particular variable. The determination of the exact crossing-time, which involves bisection algorithms and is the costlier part, is done only when the existence of this value is certain. No root search is unnecessarily performed. We have calculated the system of inequalities expressing the threshold-crossing condition, for a generic affine or linear neuronal dynamics in any dimension. The result is the conjunction of inequalities (23). It consists of two affine linear inequalities in the state-space variables, voltage and currents, and a non- linear one. The numerical implementation of this system of inequalities can be further optimized, on a case-by-case basis. In order to give a concrete implementation example of the generic inequalities (23), to show their geometrical meaning, and to give an example of optimization,in the present work we apply our procedure, step-by-step, to the 2-dimensional case of a leaky integrate-and-fire neuron with exponentially decaying post-synaptic currents (Rotter and Diesmann, 1999). The generic inequalities (23) take in this case the concrete form (44). The quantitative data in the present work are obtained by integrating these inequalities into a combined event-and-time-driven simulation framework (Morrison et al., 2007) for large-scale spiking neuronal network models as released by Bos et al. (2015). Implementation and comparison to earlier work show that: • the system of inequalities, the non-linear one in particular, can be expressed analytically in terms of the state-space variables even when the original threshold-crossing condition involves a transcendental equation – and would therefore require bisection algorithms. Compare (44) with (30); • the computationally expensive non-linear function in the system can be conveniently triangularized, speeding up the algorithm even further by testing a linear inequality first, ruling out the majority of initial states; • the new method reproduces the analytic solution of the neuron model within floating point precision. It detects all threshold crossings, in particular those that the approximate test (1) of Hanuschkin et al. (2010), misses in some ranges of mean activity, fluctuations, and synaptic-coupling strength (Figure 8); • at the default spike accuracy of 0.1 ms of the reference simulator the new method is 8 %–33 % slower than the fastest available solver with spike loss (Hanuschkin et al., 2010). It is therefore of comparable speed as embedded event driven methods (see Hanuschkin et al. (2010) Fig 5, inset). Frontiers 29 J. Krishnan et al. Perfect spike detection via time reversal In practice the method of Hanuschkin et al. (2010) rarely misses spikes for biologically realistic synaptic amplitudes. At low frequencies of afferent synaptic events, however, such missed threshold-crossings can happen. Our new scheme therefore offers an alternative for users who need guaranteed spike detection and are willing to pay a price in terms of slightly longer computation time. In state-spaces of higher dimensions it is be more difficult to derive the non-linear inequality of the system (23) in implicit form, but its set of approximating flat surfaces can still be easily calculated. In the worst-case scenario of an inequality not expressible analytically, it is still possible to construct a nested sequence of approximating flat surfaces to be tested hierarchically. Such construction only needs to be done once for any given model and detects spikes with any desired precision. Being linear, such nested inequalities likely are computationally less expensive than a bisection algorithm. The problem of detecting some sort of threshold crossing in a system of coupled first-order linear differential equations appears in many other applications and phenomena like switching, friction, and saturation (Hiebert and Shampine, 1980). For example, in an air-conditioning unit a thermostat controls the on-off state based on a certain threshold value of the room temperature (Shampine, 1994). The dynamics of this system is similar to that described in Sections 3.1–3.2. Another example is the problem of ejecting a pilot such that collision with the aircraft stabilizer is avoided. The present work use concepts from differential geometry, in particular extrusions (Bossavit, 2003) and critical points of maps between manifolds, and shows that these concepts have a readily understandable geometrical and visual meaning. The notion of extrusion has recently found applications in numerical and discretization techniques for partial and integral differential equations (Desbrun et al., 2005). The notion of critical points of a manifold mapping is ubiquitous in science: from the caustics of propagating seismic fronts, at which the seismic wave changes its phase (Romanowicz and Dziewonski, 2007, § 1.04), to the singularities between two coordinate charts in general relativity (Misner et al., 2003), which affect the accuracy of global navigation satellite systems (Coll et al., 2012; Sáez and Puchades, 2013). Indeed, the neuron model analysed in Section 3 exhibits a similarity with the dynamics of a point mass near a black hole. If the simulation timestep h is very large the curved surface separating the states that lead to a spike from those that do not acts like an event horizon in general relativity: a state evolved from the spike region can enter the no-spike region, but once there it cannot escape and will always remain a "no-spike" state. This is only true for the dynamics (6), though, with constant affine term and no resets at threshold. Inputs from other neurons lead to discontinuous changes in the affine term of the dynamics, causing a "transport" of initial states out of the event horizon, from the no-spike to the spike region. Nevertheless, maybe such similarities are more than mere coincidences. For example, the trajectory of the threshold surface of a leaky integrate-and-fire model with α-shaped post-synaptic currents (Bernard et al., 1994) can be implicitly expressed in terms of the Lambert-W function (Corless et al., 1996), as an analysis along the lines of Section 2.2 shows. This function also appears in the implicit expression of point-mass trajectories in (1 + 1)-dimensional general relativity (Mann and Ohta, 1997). It is surely worthwhile to bring the nascent field of neuronal dynamics closer to ideas and techniques from differential geometry and general relativity. This is a provisional file, not the final typeset article 30 J. Krishnan et al. Perfect spike detection via time reversal A INTERPOLATION FOR THRESHOLD-CROSSING TIME If we know that the neuron voltage V is below threshold at times t and t + h but above threshold somewhere in the interval ]t, t + h], then by continuity it must reach a maximum above threshold at a time tmax ∈ ]t, t + h]. This time can be obtained solving the equation V (tmax) = 0, with V (t) given by (26). The solution (Hanuschkin et al., 2010) is (cid:20) τ τs (cid:18) Ie I − τ − τs τ − CV τ I (cid:19)(cid:21) , tmax = − τ τs τ − τs ln (51) from which the potential V (tmax) > θ can also be easily calculated. The time tθ ∈ ]t, tmax] at which the first threshold crossing occurs, V (tθ) = θ, must lie between t and tmax, with V (t) < V (tθ) < V (tmax), and can thus be interpolated using a bisection algorithm (Press et al., 2007). B NETWORK SETUP The dynamics of the membrane potential V and synaptic current I can be described, if the input is treated stochastically and for weak synaptic couplings, by a diffusion process with equations (Fourcaud and Brunel, 2002b) τ dV dt d(RI) dt τs = −V (t) + (RI)(t), = −(RI)(t) + µ + σ √ τ ξ(t). (52) {Ji} and firing rates {ri} of the input neurons via µ = τ(cid:80) with ξ a zero-mean Gaussian process. The parameters µ and σ2 characterize the stochastic input and are the mean and variance of the total incoming synaptic current. They are related to the synaptic couplings 2ri. If we have one excitatory and one inhibitory population of input neurons, mimicked by two Poisson generators with rates rE and rI coupled to the neuron with strengths JE and JI (where J = τ C w and w is the synaptic weight of the current), and by an input current Ie, then i Jiri and σ2 = τ(cid:80) i Ji µ = Ieτ C σ2 = τ (JI + τ (JI rI + JErE) 2rE), (cid:90) θ−µ 2rI + JE √ σ + ζ(1/2)√ 2 √ τs √ τs τ (53) 1 r where r is the output firing rate of the neuron, approximated to linear order in(cid:112)τs/τ, τr is the refractory ζ(1/2)√ Vr−µ σ + 2 τ = τr + τ π ey2 [1 + erf(y)] dy, time, Vr the reset voltage, and ζ the zeta function (Abramowitz and Stegun, 1972). CONFLICT OF INTEREST STATEMENT The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest. Frontiers 31 J. Krishnan et al. Perfect spike detection via time reversal AUTHOR CONTRIBUTIONS All authors have made significant, direct, and intellectual contribution to the work. FUNDING This work was supported by DFG Grant: GSC11; Helmholtz association: VH-NG-1028 and SMHB; EU Grant 604102 (HBP); Juelich Aachen Research Alliance (JARA). ACKNOWLEDGMENTS PGLPM thanks the Forschungszentrum librarians for their always prompt and kind help, Mari & Miri for continuous encouragement and affection, Buster Keaton for filling life with awe and inspiration, and the developers and maintainers of LATEX, Emacs, AUCTEX, MiKTEX, arXiv, biorXiv, PhilSci, Python, Inkscape, Sci-Hub for making a free and unfiltered scientific exchange possible. REFERENCES Abramowitz, M. and Stegun, I. A. (eds.) (1972). Handbook of Mathematical Functions: With Formulas, Graphs, and Mathematical Tables, vol. 55 of National Bureau of Standards Applied mathematics series (Washington, D.C.: U.S. Department of Commerce), tenth printing, with corrections edn. First publ. 1964 Artin, E. (1955). Geometric Algebra, vol. 3 of Interscience tracts in pure and applied mathematics (New York: Interscience) Bernard, C., Ge, Y. C., Stockley, E., Willis, J. B., and Wheal, H. V. (1994). Synaptic integration of NMDA and non-NMDA receptors in large neuronal network models solved by means of differential equations. Biol. Cybern. 70, 267–273 Bernardo, J.-M. and Smith, A. F. (2000). Bayesian Theory. Wiley series in probability and mathematical statistics (New York: Wiley), reprint edn. First publ. 1994 Bos, H., Morrison, A., Peyser, A., Hahne, J., Helias, M., Kunkel, S., et al. (2015). Nest 2.10.0. http: //dx.doi.org/10.5281/zenodo.44222, http://www.nest-simulator.org/ Bossavit, A. (1991). Differential Geometry: for the Student of Numerical Methods in Electromagnetism. http://butler.cc.tut.fi/~bossavit/ Bossavit, A. (2002). Applied Differential Geometry (A Compendium). http://butler.cc.tut. fi/~bossavit/BackupICM/Compendium.html. First publ. 1994 Bossavit, A. (2003). Extrusion, contraction: their discretization via Whitney forms. COMPEL 22, 470–480. http://butler.cc.tut.fi/~bossavit/Papers.html Brette, R. (2007). Exact simulation of integrate-and-fire models with exponential currents. Neural Comput. 19, 2604–2609 Burke, W. L. (1987). Applied Differential Geometry (Cambridge: Cambridge University Press), reprint edn. First publ. 1985 Burke, W. L. (1995). Div, Grad, Curl Are Dead. http://count.ucsc.edu/~rmont/papers/ Burke_DivGradCurl.pdf; 'preliminary draft II'. See also http://www.ucolick.org/ ~burke/ Choquet-Bruhat, Y., DeWitt-Morette, C., and Dillard-Bleick, M. (1996). Analysis, Manifolds and Physics. Part I: Basics (Amsterdam: Elsevier), rev. ed. edn. First publ. 1977 This is a provisional file, not the final typeset article 32 J. Krishnan et al. Perfect spike detection via time reversal Coll, B., Ferrando, J. J., and Morales-Lladosa, J. A. (2012). Positioning systems in Minkowski space-time: Bifurcation problem and observational data. Phys. Rev. D 86, 084036. arXiv:1204.2241 Corless, R. M., Gonnet, G. H., Hare, D. E. G., Jeffrey, D. J., and Knuth, D. E. (1996). On the Lam- bert W function. Adv. Comput. Math. 5, 329–359. https://cs.uwaterloo.ca/research/ tr/1993/03/W.pdf Coxeter, H. S. M. (1969). Introduction to Geometry (New York: Wiley), 2 edn. First publ. 1961 de Rham, G. (1984). Differentiable Manifolds: Forms, Currents, Harmonic Forms, vol. 266 of Grundlehren der mathematischen Wissenschaften (Berlin: Springer). Transl. by F. R. Smith. First publ. in French 1955 Dehaene, S. and Changeux, J.-P. (2005). Ongoing spontaneous activity controls access to consciousness: A neuronal model for inattentional blindness. public library of science 3, 0910–0927 Desbrun, M., Hirani, A. N., Leok, M., and Marsden, J. E. (2005). Discrete exterior calculus. arXiv: math/0508341 D'Haene, M. and Schrauwen, B. (2010). Fast and exact simulation methods applied on a broad range of neuron models. Neural Comput. 22, 1468–1472 D'Haene, M., Schrauwen, B., Van Campenhout, J., and Stroobandt, D. (2009). Accelerating event-driven simulation of spiking neurons with multiple synaptic time constants. Neural Comput. 21, 1068–1099. doi:10.1162/neco.2008.02-08-707 Dodson, C. T. J. and Poston, T. (1991). Tensor Geometry: The Geometric Viewpoint and its Uses, vol. 130 of Graduate texts in mathematics (Springer), 2 edn. First publ. 1977 Ferscha, A. (1996). Parallel and distributed simulation of discrete event systems. In Parallel and Distributed Computing Handbook, ed. A. Y. Zomaya (McGraw-Hill), chap. 35. 1003–1041 Fourcaud, N. and Brunel, N. (2002a). Dynamics of the firing probability of noisy integrate-and-fire neurons. Neural Comput. 14, 2057–2110 Fourcaud, N. and Brunel, N. (2002b). Dynamics of the firing probability of noisy integrate-and-fire neurons. Neural Comp. 14, 2057–2110. http://galton.uchicago.edu/~nbrunel/pdfs/ fourcaud02.pdf Fujimoto, R. M. (2000). Parallel and distributed simulation systems (New York: Wiley) Gewaltig, M.-O. and Diesmann, M. (2007). NEST (NEural Simulation Tool). Scholarpedia 2, 1430 Good, I. J. (1966). How to estimate probabilities. J. Inst. Maths. Applics 2, 364–383 Hansel, D., Mato, G., Meunier, C., and Neltner, L. (1998). On numerical simulations of integrate-and-fire neural networks. Neural Comput. 10, 467–483 Hanuschkin, A., Kunkel, S., Helias, M., Morrison, A., and Diesmann, M. (2010). A general and efficient method for incorporating precise spike times in globally time-driven simulations. Front. Neuroinform. 4, 113 Hiebert, K. L. and Shampine, L. F. (1980). (Albuquerque, USA: Sandia Laboratories) Implicitly Defined Output Points for Solutions of ODEs Johnson, W. E. (1932). Probability: The deductive and inductive problems. Mind 41, 409–423. With some notes and an appendix by R. B. Braithwaite Krishnan, J., Porta Mana, P., Helias, M., Diesmann, M., and Di Napoli, E. (2016). Perfect detection of spikes via time-reversal. In Proceedings of the Bernstein Conference. doi:http://dx.doi.org/10.12751/ nncn.bc2016.0152 Kunkel, S., Helias, M., Diesmann, M., and Morrison, A. (2011). Fail-safe detection of threshold crossings of linear integrate-and-fire neuron models in time-driven simulations. BMC Neuroscience 12, P229 Frontiers 33 J. Krishnan et al. Perfect spike detection via time reversal Mann, R. B. and Ohta, T. (1997). Exact solution for the metric and the motion of two bodies in (1 + 1)-dimensional gravity. Phys. Rev. D 55, 4723–4747. arXiv:gr-qc/9611008 Marsden, J. E. and Ratiu, T. (2007). Manifolds, Tensor Analysis, and Applications, vol. 75 of Applied mathematical sciences (New York: Springer), 3 edn. http://www.esm.vt.edu/~sdross/pub/ books/. Written with the collaboration of Ralph Abraham. First publ. 1983 Mattia, M. and Del Giudice, P. (2000). Efficient event-driven simulation of large networks of spiking neurons and dynamical synapses. Neural Comput. 12, 2305–2329 Misner, C. W., Thorne, K. S., and Wheeler, J. A. (2003). Gravitation (New York: W. H. Freeman and Company), 25th printing edn. First publ. 1970 Morrison, A. and Diesmann, M. (2008). Maintaining causality in discrete time neuronal network simu- lations. In Lectures in Supercomputational Neuroscience: Dynamics in Complex Brain Networks, eds. P. beim Graben, C. Zhou, M. Thiel, and J. Kurths (Springer), Understanding Complex Systems. 267–278 Morrison, A., Straube, S., Plesser, H. E., and Diesmann, M. (2007). Exact subthreshold integration with continuous spike times in discrete time neural network simulations. Neural Comput. 19, 47–79 Nomizu, K. and Sasaki, T. (1994). Affine Differential Geometry: Geometry of Affine Immersions, vol. 111 of Cambridge tracts in mathematics (Cambridge: Cambridge University Press) Porta Mana, P. G. L. (2011). Notes on affine and convex spaces. arXiv:1104.0032 Press, W. H., Teukolsky, S. A., Vetterling, W. T., and Flannery, B. P. (2007). Numerical Recipes: The Art of Scientific Computing (Cambridge: Cambridge University Press), 3 edn. First publ. 1988 Ramanan, S. (2005). Global Calculus, vol. 65 of Graduate studies in mathematics (Providence, USA: American Mathematical Society) Rauch, A., La Camera, G., Lüscher, H., Senn, W., and Fusi, S. (2003). Neocortical pyramidal cells respond as integrate-and-fire neurons to in vivo like input currents. Journal of Neurophysiology 90, 1598–1612 Rockafellar, R. T. (1972). Convex Analysis, vol. 28 of Princeton mathematical series (Princeton: Princeton University Press). First publ. 1970 Romanowicz, B. and Dziewonski, A. (eds.) (2007). Seismology and the Structure of the Earth, vol. 1 of Treatise on Geophysics (Amsterdam: Elsevier), 1 edn. Rotter, S. and Diesmann, M. (1999). Exact digital simulation of time-invariant linear systems with applications to neuronal modeling. Biol. Cybern. 81, 381–402 Sáez, D. and Puchades, N. (2013). Relativistic positioning systems: Numerical simulations. Acta Futura 7, 103–110. arXiv:1404.1000 Schouten, J. A. (1989). Tensor Analysis for Physicists. Dover books on physics and chemistry (New York: Dover), corr. second ed. edn. First publ. 1951 Schücker, J., Diesmann, M., and Helias, M. (2015). Modulated escape from a metastable state driven by colored noise. Phys. Rev. E 92, 052119. arXiv:1411.0432 Shampine, L. F. (1994). Numerical Solution of Ordinary Differential Equations (New York: Chapman & Hall/CRC) Shelley, M. J. and Tao, L. (2001). Efficient and accurate time-stepping schemes for integrate-and-fire neuronal networks. Journal of Computational Neuroscience 11, 111–119 Simon, U., Schwenk-Schellschmidt, A., and Viesel, H. (1992). Introduction to the Affine Differential Geometry of Hypersurfaces. Lecture notes of the Science University of Tokyo (Tokyo: Science University of Tokyo) Spivak, M. (1999). A Comprehensive Introduction to Differential Geometry. Vol. One (Houston, USA: Publish or Perish), 3 edn. First publ. 1970 This is a provisional file, not the final typeset article 34 J. Krishnan et al. Perfect spike detection via time reversal van Elburg, R. A. J. and van Ooyen, A. (2009). Generalization of the event-based Carnevale-Hines integration scheme for integrate-and-fire models. Neural Computation 21, 1913–1930. doi:10.1162/ neco.2009.07-08-815 Zabell, S. L. (1982). W. E. Johnson's "sufficientness" postulate. Ann. Stat. 10, 1090–1099. Repr. in (Zabell, 2005, pp. 84–95) Zabell, S. L. (2005). Symmetry and Its Discontents: Essays on the History of Inductive Probability. Cambridge studies in probability, induction, and decision theory (Cambridge: Cambridge University Press) Zeigler, B. P., Praehofer, H., and Kim, T. G. (2000). Theory of Modeling and Simulation: Integrating Discrete Event and Continuous Complex Dynamic Systems (Amsterdam: Academic Press), 2 edn. Zheng, G., Tonnelier, A., and Martinez, D. (2009). Voltage-stepping schemes for the simulation of spiking neural networks. Journal of Computational Neuroscience 26, 409–23 Frontiers 35
1704.02342
1
1704
2017-04-07T18:33:59
Sleep Paralysis: phenomenology, neurophysiology and treatment
[ "q-bio.NC" ]
Sleep paralysis is an experience of being temporarily unable to move or talk during the transitional periods between sleep and wakefulness: at sleep onset or upon awakening. Feeling of paralysis may be accompanied by a variety of vivid and intense sensory experiences, including mentation in visual, auditory, and tactile modalities, as well as a distinct feeling of presence. This chapter discusses a variety of sleep paralysis experiences from the perspective of enactive cognition and cultural neurophenomenology. Current knowledge of neurophysiology and associated conditions is presented, and some techniques for coping with sleep paralysis are proposed. As an experience characterized by a hybrid state of dreaming and waking, sleep paralysis offers a unique window into phenomenology of spontaneous thought in sleep.
q-bio.NC
q-bio
Elizaveta Solomonova1,2 Abstract Sleep Paralysis: phenomenology, neurophysiology and treatment Clinical Conditions. Fox, K & Christoff, K. Eds. 1Université de Montréal, Individualized program (Cognitive Neuroscience & Philosophy). 2Center for Advanced Research in Sleep Medicine, Dream and Nightmare Laboratory, Montreal, Canada To appear in: The Oxford Handbook of Spontaneous Thought: Mind-Wandering, Creativity, Dreaming, and Sleep paralysis is an experience of being temporarily unable to move or talk during the transitional periods between sleep and wakefulness: at sleep onset or upon awakening. Feeling of paralysis may be accompanied by a variety of vivid and intense sensory experiences, including mentation in visual, auditory, and tactile modalities, as well as a distinct feeling of presence. This chapter discusses a variety of sleep paralysis experiences from the perspective of enactive cognition and cultural neurophenomenology. Current knowledge of neurophysiology and associated conditions is presented, and some techniques for coping with sleep paralysis are proposed. As an experience characterized by a hybrid state of dreaming and waking, sleep paralysis offers a unique window into phenomenology of spontaneous thought in sleep. Introduction "I had a few terrifying experiences a few years ago. I awoke in the middle of the night. I was sleeping on my back, and couldn't move, but I had the sensation I could see around my room. There was a terrifying figure looming over me. Almost pressing on me. The best way I could describe it was that it was made of shadows. A deep rumbling or buzzing sound was present. It felt like I was in the presence of evil... Which sounds so strange to say!" (31 year old man, USA) Sleep paralysis (SP) is a transient and generally benign phenomenon occurring at sleep onset or upon awakening. Classified as a rapid-eye-movement (REM) sleep-related parasomnia, SP represents a psychophysiological state characterized simultaneously by qualities of both sleep and wakefulness, wherein the experiencer can open her eyes (Hishikawa & Kaneko, 1965), can be aware of her physical environment but is unable to move and may start seeing, hearing, feeling or sensing something. While documented instances of SP seem to be very consistent across cultures , SP's lived qualities, phenomenology, and interpretation as a meaningful experience varies depending on the cultural and religious background. Rooting SP in a particular belief system may either help the experiencer recognize that SP is common and transient, or amplify negative qualities of SP by giving more concrete shape to an already terrifying experience of a supernatural assault. The emotional experience of SP is often one of fear, terror and panic. Threatening presences, the vulnerability of being in a paralyzed state, uncontrollable visions - all these elements contribute to intense, predominantly dysphoric, negative affect. Some qualities of spontaneous thought associated with felt presence during SP can be seen as paranoid (Cheyne & Girard, 2007), spatial, or interpersonal/social (Nielsen, 2007, Solomonova, 2008). The vast majority of SP experiences are associated with intense feelings of realism, and are most often characterized by fear and distress, which may carry-over into wakefulness and create a vicious cycle of negative emotional association with sleep, including aversion to going to bed, and even, in extreme cases, can result in symptoms reminiscent of post-traumatic stress disorder (McNally and Clancy, 2005). Yet, some SP experiences are described in positive terms, especially vestibulo-motor phenomena that include out-of-body experiences (OBE), or sensations of flying or floating. While the intuitive and immediate reaction to SP is typically negative, as will be discussed in the section on practical considerations, there are numerous reports of neutral/positive SP. Furthermore, there is a possibility of harnessing the power and potential of the dissociative/overlapping state in order to take active charge of 1 one's experience and to use the opportunity presented by the simultaneity of waking and sleeping cognition - a potential for entering into a lucid dream state or for contemplative self-observation. Neurophysiologically, SP is currently understood as a state dissociation or a state overlap between REM sleep and wakefulness (American Academy of Sleep Medicine, 2014). During SP one can open her eyes, look around the room, become aware of her environment and simultaneously experience REM sleep-related paralysis (muscle atonia) as well as intense and realistic imagery1 of all sensory modalities - a nightmare spilling into the real world. Normally, during REM sleep, skeletal muscle atonia blocks most motor output, effectively preventing the sleeper from acting out her dreams (Peever, Luppi, & Montplaisir, 2014). SP can also occur in the context of narcolepsy (Sharpless & Barber, 2011; Terzaghi, et al, 2012), but the majority of those who experience SP report it in its isolated form (often referred to as Isolated Sleep Paralysis), without known medical or neurological association. In current medical and neuroscientific literature, SP is discussed in terms of its presentation and negative factors: SP-associated mentation is generally seen as a non-desirable effect of REM sleep intrusion into waking. In this chapter I propose that situating SP experiences as a dream phenomenon within the framework of embodied mind and enactive cognitive science, including contemplative approaches to consciousness, is an alternative that accounts for the phenomenology of SP as a lived experience, allows for rich and detailed cultural framing of the experience, and offers avenues for a cross-cultural social neurophenomenology of SP. First, I will discuss the phenomenology and neurophysiology of SP experiences. I will present SP in general, without distinguishing between its isolated and narcolepsy-related form, unless such a separation is warranted. I will start by presenting the current state of knowledge of SP prevalence, as well as the varieties of imagery accompanying SP and their cultural significance. I will then discuss SP in terms of a REM sleep parasomnia and outline its precipitating and enabling factors, as well as sleep and dream characteristics of those who experience SP. Finally, I will examine SP in light of various cultural and shared practices, including preventative measures, and practices aimed at interrupting and transforming SP experiences. The experiential examples of SP used in the present chapter are all derived from an Internet-based study of SP (Solomonova et al, 2008). Our research group collected 193 responses from people with recurrent SP experiences, using a modified version of the Waterloo Unusual Sleep Experiences Questionnaire (Cheyne, Newby-Clark, & Rueffer, 1999). Our participants were recruited online, using word of mouth, and via advertising on SP-related forums, information and support. Idiopathic SP (SP not associated with narcolepsy, and without known cause) is a benign and transient parasomnia (Howell, 2012), occurring during transitions between wake and sleep: at sleep onset or upon awakening. The Diagnostic and Statistical Manual of Mental Disorders (DSM-5) classifies isolated SP accompanied by fearful mentation as an instance of a nightmare disorder (American Psychiatric Association, 2013). During an episode of SP, characteristics of REM sleep intrude upon seemingly awake consciousness: thus the person experiencing SP, while having an impression of being awake and aware of her environment, is unable to initiate voluntary movements (i.e., experiences REM sleep muscle atonia/paralysis), and may also experience intense and realistic sensations in any sensory modality - REM sleep-related mentation (American Academy of Sleep Medicine, 2014). SP should be distinguished from night terrors – early night awakenings with feelings of panic/terror, typically associated with somnambulism-spectrum arousal from slow wave (Stages 3 and 4) non-REM (NREM) sleep. Night terrors are characterized by sudden awakening in an agitated state, anxiety, body motility and general amnesia with regards to underlying cognitive experience (Szelenberger, Niemcewicz & Dabrowska, 2009). Prevalence estimates of SP range widely, and may depend on geographic and cultural factors. A systematic review of SP prevalence has revealed that studies report SP lifetime prevalence from as low as 1.5% to possibly 100% in the general population (Sharpless & Barber, 2011). The authors indicated that 1 In this chapter I will use the terms 'imagery' and 'mentation' interchangeably to refer to visual, auditory, somatosensory and even social experiences during SP. The term 'imagery' here is therefore not restricted to the visual domain. I prefer 'imagery' and 'mentation' to 'hallucination' in order to emphasize the dream-like process of spontaneous imagination that takes place during SP, and to de-emphasize the association with delusional thought and pathologies, associated with the term 'hallucination'. Definitions and prevalence 2 The 4EA cognition and oneiric mentation about one in five individuals may have experienced SP at least once in their lifetime (of 36,533 persons in their review). Prevalence estimation of SP is difficult due to numerous factors such as ethnicity and cultural background, including variable familiarity with the phenomenon and in the wording of questions (Fukuda, 1993). For example, in a cross-cultural study, Fukuda and colleagues (2000) reported that while it is unclear whether SP is equally prevalent in Canada and Japan, the lack of familiarity and of a normative cultural framework for SP in Canada may contribute to the fact that many Canadian, but not Japanese, respondents qualified SP as a kind of a dream, and would not have, therefore, readily recognized SP in a prevalence study. An additional reason for under-diagnosis of SP in the West may be the fact that those who experience SP may be misdiagnosed as having psychiatric disturbances (Hufford, 2005). The developmental trajectory of SP is traditionally associated with an onset during adolescence, which may indicate a process associated with sleep architecture maturation (Wing, Lee & Chen, 1994). However, in one study of older adults, a bimodal onset pattern was reported, with a second pattern of onset of SP episodes after the age of 60 years old (Wing et al, 1999), suggesting a possibility that SP may have a variety of onset conditions. Neurologically and phenomenologically SP is situated on the REM-sleep based dream/nightmare continuum. In this chapter, I will approach SP experiences as a variant of intensified or disturbed dreaming, and will situate them within a framework of embodiment and enactivism. Recent years have seen the development of a paradigm shift from a strictly neurocentric view of the mind, a position that can be stated as "embrained" (Morris, 2010), to a diverse family of approaches that consider the mind embodied (Varela, Thompson & Rosch, 1991; Gallagher, 2005), enactive (Noe, 2004; Thompson, 2005; Stewart, Gapenne & Di Paolo, 2010), extended into and embedded in the physical and social world (Clark & Chalmers, 1998; Menary, 2010), and affective (Colombetti, 2013; Pessoa, 2013). While these approaches are in many respects quite different, they have sometimes been labeled as 4EA (embodied, embedded, extended, enactive, affective) cognition, with a common theme of offering a robust alternative to computational, connectionist, and neuro-reductionist views of the mind (Wheeler, 2005; Protevi, 2012). These theories attempt to situate cognition, brain activity, and psychophysiology within the larger contexts of lived subjective experience, by emphasizing the roles of developmental sensorimotor attunement to the world, as well as of the active and motivated processes of perception and sense-making, the importance of the social and cultural milieu, and the role of emotion and affect. Sleep and dreaming phenomena have been only rarely addressed by 4EA theorists (with the exception of Thompson 2014, 2015a,b), and the prevailing view of the sleeping mind today situates sleep mentation as being firmly constrained within the brain (Rechtschaffen, 1978; Hobson, Pace-Schott & Stickgold, 2000; Revonsuo, 2006). As Revonsuo states: "The conscious experiences we have during dreaming are isolated from behavioral and perceptual interactions with the environment, which refutes any theory that states that organism-environment interaction or other external relationships are constitutive of the existence of consciousness" (Revonsuo et al, 2015: 3). Alternatively, situating dream mentation within a framework of 4EA approaches implies that the dreaming subject is not entirely isolated or disconnected from environmental and somatic stimuli, and that her experiential self retains affective, social, sensorimotor and sense-making qualities. Dreaming then is not passively lived as a purely mental simulation (Revonsuo et al, 2015), but can be seen as a process of active imagination (Thompson, 2014) rooted in the dreamer's physical, social and affective world (Solomonova & Sha, 2016). I propose that SP experiences, by virtue of their special kind of overlap between and simultaneous presence of both waking and dreaming cognition, are perfect candidates for neurophenomenological research on spontaneous thought in sleep, which would help illuminate particular qualities of dreaming cognition that may otherwise be inaccessible to reflective consciousness upon awakening from a dream. 3 Phenomenology of sleep paralysis experiences Figure 1. A representation of sleep paralysis experience2. The sleeper is awakened suddenly and sees a menacing shadowy creature on top of him. He experiences the sensation of being pushed into the bed, while the bed itself is swirling in a sort of a tornado. The two faces of the dreamer represent the "double" consciousness during sleep paralysis: he is simultaneously terrified of the supernatural attacker and also knows that if he does not resist the experience and allows himself to drift back into sleep he may have a lucid dream (this lucid consciousness is represented as a sleeping face with a colorful brain, denoting vibrant possibilities of lucidity. Artist: Benjamen Samaha, Montreal, Canada 2016. Reproduced with artist's permission. In addition to transient experience of muscle paralysis, the most dramatic quality of SP is its sensory content, characterized by vivid, intrusive audio-visual and somatosensory imagery. The experience of SP can be extremely realistic, have a quasi-perceptual and wake-like quality, and may be accompanied by tactile and kinesthetic sensations. Reflective thought processes, self-awareness and metacognitive abilities seem to be relatively preserved during SP experiences, and people who have had multiple SP experiences may develop a "feel" for recognizing SP imagery. SP-associated experiences are typically referred to as hallucinations (hypnagogic, when occurring at sleep onset, or hypnopompic, when happening upon awakening), since these occur during otherwise seemingly awake consciousness (Liddon, 1967; American Academy of Sleep Medicine, 2014). This entails that a person who experiences SP sees something that is not there, something that is distorted or false. Such a view presupposes that during SP one is effectively awake and is misinterpreting her experience. Another way of looking at SP is to situate it within the spectrum of dream mentation and dreaming imagination. While dreaming too has been seen as a delusional/hallucinatory activity, an alternative view, in line with embodied mind theories and enactive approach, has also been proposed: "When you hallucinate, you seem to perceive 2 Excerpt from the dreamer's account: "The transition between wake and sleep is a crucial moment to enter into the world of dreams. During this transition on countless occasions I would awaken suddenly not being able to move. During this experience it seems that the very essence of fear permeates my consciousness. Eeriness goes through my soul, freezes my blood and interrupts all substantial notion of my being. No words can describe that visceral sensation, and in parallel, no words can come out of my mouth. Aware of the lack of muscle tonus, I try to escape this inevitable Machiavellian black beast, that materializes in my head before my eyes and on my chest, slowing down my breathing. Moreover, my senses are grabbed by an impression of fighting a hurricane that may drag me out of my body. This cyclone, that has a black hole in lieu of an eye, forces me to fight it, and this fight seems crucial to my survival (…) In addition, there are auditory experiences, an amalgam of petrifying words and vibrations that feel like sudden gusts of wind in the eardrums. All this happens when by body feels like a statue, without a possibility of screaming. (…) On occasion, with determination and lucidity, I can have power over this swirl of stillness. (…) I take back some control of my imaginary hands, and then I hold them out to Morpheus for a dazzling and colorful dance in a deep and enlightened night". (Male SP sufferer, also diagnosed with narcolepsy. Montreal, Canada. Account translated from French). 4 Kinds of sleep paralysis experience what is not there. (…) When you imagine, you evoke something absent and make it mentally present to your attention" (Thompson, 2014: 179). In this chapter I adopt this latter view and will refer to SP experiences as variant of spontaneous thought/mentation or mental imagery, rather than hallucinations or delusions. A factor analysis by Cheyne and colleagues (1999) showed that SP mentation typically falls into three general categories. The first category is Intruder, and it is characterized by a felt presence, fear, as well as auditory and visual imagery. The person who experiences SP feels that someone is in the house or in their room. This experience is sometimes accompanied by seeing or hearing someone or something sentient move around the house. The second category of SP experiences is known as Incubus, in which the felt presence is interpreted as a supernatural assault and is often accompanied by a sensation of shallow breathing, a feeling of being smothered, pressure on the chest, or pain. In this case, the sleeper often sees and feels the maleficent being on top of her. The third category, Unusual Bodily Experiences, appears to be a separate, less well-known, and a qualitatively different kind of SP experience: these are often described as positive events, such as sensations of floating, out-of-body experiences (OBEs), and feelings of bliss. Both Intruder and Incubus categories typically include the experience of felt presence – a distinct sensation that someone sentient is in the immediate vicinity of the sleeper (Cheyne, 2005). Most literature on SP focuses almost exclusively on the first two kinds, Intruder and Incubus, possibly due to their particularly intensified felt presence imagery, which contributes to distressing SP experiences (Solomonova et al, 2008; Cheyne, 2013). However, neutral and positive instances of SP have also been described, and the third category, Unusual Body Experiences, or vestibulo-motor experiences, is often characterized by pleasant sensations and a spirit of exploration, accompanying sensations of flying, out-of- body experiences, or autoscopy (observation of one's own body from an unusual/novel point of view) (Brugger, Regard & Landis, 1997). figure is standing over the bed staring down at me, or pacing back and forth." (22 year old, gender not reported, USA). Among all SP experiences, felt presence, the distinct sensation that another sentient being, human or not, is present in the extracorporeal space of the experiencer, is arguably the most salient, terrifying, and rich.. Felt presence is consistently reported as the most common SP-associated experience – about 80% of episodes (Cheyne et al, 1999), which produces most fear and SP-related state of distress (Solomonova et al, 2008). One salient feature of felt presence experiences during SP is the fact that it is a distinct sensation, and may occur in the absence of visual, auditory, or tactile imagery. Felt presence experiences during SP have been classified as a paranoid delusion (Cheyne & Girard, 2007), an expression of spatial social imagery (Nielsen, 2007), and as a variant of basic intersubjective experience of the world (Solomonova, Frantova, & Nielsen, 2010). Felt presence experiences are often interpreted within the cultural framework available to the experiencer (see the following section on the cultural neurophenomenology of SP), but some basic characteristics seem to be common across cultures and ages (Cheyne, 2001): 1) felt presence often manifests from ambiguous stimuli: it is often described as "shadowy", and its physical characteristics are often unclear; 2) the experiencer may report a distinct sensation of being watched, and that the presence has some intentions towards the dreamer; these range from some vague interest to full-blown assault; 3) felt presence is usually accompanied by intense emotions (often fear when the presence is interpreted as threatening), sometimes to the point of a distinct feeling of dread, imminent death, or being in the presence of evil. Positive emotions, however, are also possible, especially when the experience is understood as visitations by deceased relatives or visions of the divine. Consider the following examples of felt presence experiences of the Intruder type: A 26-year-old man from the United States reports: "It felt as if someone was watching me but silently standing behind me". In this example the presence is felt in a distinct and clear way, but not seen or heard, yet the experiencer knows where in space the presence is located. Similarly, a 29-year-old woman from USA regularly experiences the "Just before going to sleep or if awoken suddenly I feel as though a presence, usually a dark shadow Felt Presence Intruder 5 Incubus malevolent presence without ever seeing it: "…feeling of evil that is watching or monitoring; never able to actually see this "evil entity". Even in the absence of direct visual, auditory or tactile imagery she feels that she is observed and that the presence is "evil". The ambiguous qualities of the physical attributes of SP visitors can be illustrated by the following two examples. A 39-year-old man from USA writes: "The "presence" is a tall black/darkest grey shadow of a human form without any features. It stands in the doorway to my bedroom waiting to be "noticed". Likewise, a 30-year-old woman experienced various ways in which the presence was manifesting during her SP attacks: "Once it seemed a shadow was leaving the room. One other time the shadow seemed to have "wild" hair or if it doesn't have hair at all, it looked as some sort of black something". The Incubus experience happens when the Intruder physically oppresses the sleeper, sometimes in a rather dramatic way. In words of a 52-year-old man from the United States: "My worst experience was being choked by a man who burst into my bedroom. The experience was so real and frightening that I was very afraid of my SP for many months after." The Incubus takes many forms, including human, supernatural and more rarely, animal: "I often hallucinate creatures like large cats - lions or tigers, … wrapping themselves firmly around me and crushing my body" writes a 20-year-old woman from England. Some of the most dramatic and potentially traumatic SP Incubus experiences are instances that are lived as sexual assault or alien abduction. Consider the following example, reported by a 40-year-old man from the United States: "When it is a "Dark Man" episode, he most likely touches me. Either by laying across my body, in a sexual way or in the beginning, he would grab me and drag me. I always felt that if I let go, he would pull me out of my body". Similarly, a 29-year-old woman from Spain describes her distressing SP experience: "…extreme terror, the feeling that air is dense and darker, that shadows boil and take shape… I hear some low tone noises, voices, tactile feeling of grabbing, of naked cold skin, and, very rarely, a presence. Very dark with round eyes, spider-like fingers, that laughs, messes up the bed, and makes me feel terror, with some sexual approaches…". In a study linking reports of space alien abduction to SP episodes McNally and Clancy present this case: "…female abductee… was completely paralyzed, and felt electrical vibrations throughout her body. She was sweating, struggling to breathe, and felt her heart pounding in terror. When she opened her eyes, she saw an insect-like alien being on top of her bed" (McNally & Clancy, 2005: 116). While most easily recognizable and most commonly documented cases of felt presence during SP have to do with a threatening and ominous "visitor", some evidence suggests, however, that the presence is not always understood as hostile. Such experiences include perception of friends and family; visitation from deceased relatives or benevolent spirits; and erotic encounters where the sense of presence is comforting. A 20-year-old SP sufferer from the United States writes: "Once or twice I have thought that my friend or roommate was standing over me. I was confused but not afraid." Similarly, encountering deceased family members in visions or in dreams can be experienced as a positive spiritual event, and possibly play a healing role in processes of bereavement (Garfield, 1996; Belicki et al, 2003). Finally, another rare kind of SP-related felt presence episode involves first an experience of "someone there", and then a doubling of the dreamer's own body, a self-projection into the extracorporeal space. Some individuals report that the felt presence entities are becoming an externalized view of themselves: "Sometimes I feel that the presence is myself, that I can watch myself", reports a 21-year-old man from Jamaica; "I switch to another world and I myself become a presence", writes a 19-year-old man from Russia. Most (if not all) SP episodes are defined by an altered experience of the body. These include simple experience of muscle paralysis; sensations associated with supernatural assault, including touch, pressure on the chest, or even choking; feelings of unusual vibrations or falling into a vortex; and out-of-body experiences, including flying, falling, or moving around one's house. One of the most salient features of SP is the REM sleep-related muscle atonia. The inability to move is a striking and unusual experience for most individuals, and the mismatch between sensing the body and the Positive felt presence experiences and doubling Body experiences in sleep paralysis 6 loss of voluntary control over the body's movements may contribute to a range of somatosensory experiences. As discussed above, some of the most intense SP episodes may involve a feeling of being assaulted or touched by a supernatural entity. For instance, a 34-year-old man from USA describes the following experience: "Felt my arms pinned across my chest in a strait jacket hold, felt hands on my chest pinning me against a wall". Perception of not being able to fully breathe, often accompanied by feeling of pressure on the chest, may be prevalent in as much as 57% of SP episodes (Sharpless et al, 2010). Although most accounts of and research on SP experience have centered on paralysis accompanied by terrifying mentation and by felt presence, not all SP experiences are characterized by imagery and many are simply experiences of transient body paralysis during the transition between sleep and wakefulness, without any other accompanying mental activity (American Academy of Sleep Medicine, 2001). Additionally, SP episodes may be predominantly somatosensory in nature: Cheyne (1999) characterizes these experiences as Vestibulo-Motor mentation. Autoscopy, out-of-body experiences, vibrations, floating, falling and body doubling experiences (Cheyne, 2002) are all possible within the SP framework due to its reliance on dream-supporting REM sleep mechanisms. During a dream, especially a lucid dream (wherein one is aware of the fact that she is dreaming), it is possible to have simultaneous experience of one's dream body and real body at the same time. Thompson (2014) distinguishes between the dreaming self (I the dreamer) and the dream ego (I as dreamt) as two coinciding modes of self-experience, which may sometimes be experienced in parallel. The dreaming self is the sleeping self, it is the "I" of the waking life, now engaging in the practice of sleep and dreaming. The dream ego, on the other hand, is the experiential self, immersed in the dream scenario. The I as dreamt is the temporary "I" that takes on the first-person perspective as a subject (and sometimes an object) of the dream world. Seen from this point of view, SP episodes may represent an intense experience of the dreaming ego, lacking a dream body and temporarily "stuck" within her immobilized sleeping body of the dreaming self/I the dreamer while experiencing dream-like mentation. This feeling of being stuck, coupled with awareness of the overlap between states of vigilance, may then transform itself into a situation of perceptual doubling of body imagery. Contrary to most SP episodes with a felt presence component, some bodily experiences are described in quite positive terms. For instance, a 20-year-old woman from England describes this characteristic of her typical SP episode: "Generally, the experiences start with a low, pleasant vibration that moves through my body in defined waves, from the feet up. I feel them most strongly in the throat and in my eardrums". Out-of-body experiences are also relatively common in SP – as much as 39% of SP experiencers have had one at some point (Cheyne, 2002). A 39-year-old woman from the United States writes: "I floated out of my bed into the kitchen. But, as I floated over my bed, I saw like this beast figure crouched over on the front of my bed. I floated over it down to the kitchen. That is where I saw this beautiful kaleidoscope-like leaves. They were so vibrant … I then floated back to my room into my body". In this example there is a combination of various SP characteristics: dream-like mentation superimposed onto the environment, a nocturnal visitor, and an altered sense of the body. SP experiences are also sometimes accompanied by false awakenings-dreams where one has a vivid and realistic feeling of waking up in their own bed and engaging in usual activities only to realize that they are still asleep (Buzzi, 2011). While false awakenings are typically characterized as dream experiences, their phenomenology in terms of realism and possible state overlap is to a degree similar to SP. In Cheyne's report (2002), 58% of people who experience SP also experienced false awakenings at least occasionally. Additionally, false awakenings are often associated with feelings of dread, anxiety, and oppression (Green & McCreery, 1994; Nielsen and Zadra, 2011), similarly to SP. The following two examples from our Internet- based sample illustrate such cases: a 24-year old man from the United States reports: "… sometimes I think I have moved... sometimes even gotten up and walked around only to find that I never got up at all." In a similar vein, a 21-year old man from Jamaica describes his experience: "I will wake up into another dream inside my bedroom and think I am awake and realize I am still sleeping minutes later and the same procedure repeats several times". communicate with those around me. (23-year-old woman, USA) …Extreme anxiety and fear, mind is awake, but body is asleep. I feel as though I am trapped and cannot Emotions 7 The most prevalent emotion associated with SP experiences is fear. Indeed, the most natural reaction to waking up unable to move is panic, and the sensation of constricted breathing (consistent with REM sleep physiology) may increase the state of distress. Sharpless and colleagues (2010) introduced the term fearful isolated sleep paralysis to denote SP experiences characterized by an intense state of distress. As much as 90% of reported SP episodes are described as fearful (Cheyne, Rueffer, & Newby-Clark, 1999). Similarly, in an Irish University students' sample, fear was found to be the most prevalent emotion, with 82% of respondents stating that they have experienced fear at some point during a SP episode (O'Hanlon, Murphy, & Di Blasi, 2011). Moreover, nightmare frequency was previously reported as a predictor of SP occurrence (Liskova et al, 2016). This data suggests that SP, or at least the Fearful form of SP, can be seen as an intensified form of a nightmare: a recent study by Robert and Zadra (2014) reported that about 65% of nightmares and 45% of bad dreams are characterized by fear. One approach to classify the affective and personal impact of SP experiences is to assess not only frequency or intensity of SP episodes, but also distress associated with SP experiences (Solomonova et al, 2008; Cheyne & Pennycook, 2013). To what extent is the individual affected by SP? To what extent do negative emotions carry-over from an SP episode into waking life? Do SP experiences promote a negative relationship with sleep? These questions have been successfully examined in previous research on nightmares (Belicki, 1992; Blagrove, Farmer & Williams, 2004), showing that the individual impact of negative and intense dream experiences depends more on a trait-like reactivity, sometimes referred to as affect distress (Nielsen & Levine, 2007). This trait is thought to represent a general dysfunction of affect regulation network, and it has been shown to be a better measure of how much nightmares influence waking life emotional well-being than frequency or intensity of self-reported nightmare occurrence. Furthermore, affect distress mediates reactivity, negative interpretation and degree of negative reaction to nightmares (Belicki, 1992; Levin & Fireman, 2002). According to Nielsen and Levine (2007; Levine & Nielsen, 2009), dreaming helps regulate emotional memory consolidation and emotional reactivity via fear extinction. Nightmares, therefore, represent a case of problematic/dysfunctional processes of fear extinction. In combination with other factors, affect distress is likely to play a role in formation, experience and interpretation of SP. Positive emotions associated with SP are much less studied, and it is not possible to accurately estimate their prevalence. One possible reason for this is lack of appropriate screening (SP is often diagnosed as an unpleasant phenomenon) and lack of medical/psychiatric concern: patients are not very likely to describe such experiences to their health practitioner, since they are not bothered by them. In addition, the current diagnostic criteria for a recurrent isolated SP as listed in the latest edition of the International Classification of Sleep Disorders – 3d edition (American Academy of Sleep Medicine, 2014), include that the episodes must cause "clinically significant distress including bedtime anxiety or fear of sleep". Such a provision would effectively exclude all possible positive and non-distressing SP phenomena from investigation and/or diagnosis. Nonetheless, in a web-based SP study Cheyne (2002) reports that in addition to anger (30% of respondents) and sadness (23%), bliss (17%) and erotic sensations (17%) are also sometimes present in SP. Felt presence is the most prevalent, the most emotionally disturbing, and the most salient SP-related experience. Therefore, it is unsurprising that most visual and auditory mentation during SP usually has something to do with these unwelcome visitors. The entities, however, while felt in a very distinctive and concrete way, are often described visually as rather general and vague shadowy beings. Visual experiences are reported to occur in 54%– 56% and auditory experiences in 55%-60% (Solomonova et al, 2008; Cheyne, 2002) of SP sufferers. SP may be accompanied by auditory experiences, ranging from abstract and mechanical sounds, such as electric sounds and sounds of buzzing, to vivid auditory imagery, consistent with SP experience of an Intruder or an Incubus. Sounds of footsteps and of voices are often reported (Cheyne, Rueffer & Newby-Clark, 1999; Cheyne, 2002; Solomonova et al, 2008). Visual and auditory experiences 8 Cultural grounding of SP Figure 2. Henry Fuseli. The Nightmare. 1781. Detroit Institute of Art. Public Domain image: https://commons.wikimedia.org/w/index.php?curid=15453518 While sleep paralysis is a lesser-known sleep phenomenon in the West, it is quite prevalent and is well-described in many other cultures. Due to the lack of general awareness of SP in the West, it is rarely discussed in the context of family medicine or psychology. Cross-cultural work on SP revealed that it is rooted in a variety of religious beliefs and cultural schemas, including interpretations of the experience and techniques to engage with the nocturnal visitors. Some of the common qualities of SP across cultures (Adler, 2011) include: 1) sensation of being awake; 2) perception of the environment; 3) paralysis; 4) feeling of fear and dread; 5) felt presence; 6) chest pressure/breathing difficulties; 7) supine position; 6) various unusual body sensations. These apparently culturally-invariant qualities of SP-related occurrences of the experience of a supernatural attack have been at the center of the phenomenological and cross-cultural cognitive research on SP. Figure 2 is a reproduction of an eighteenth-century work by Henri Fuseli entitled The Nightmare. It represents a sleeping woman in a supine position being oppressed by a maleficent creature sitting on her chest and with an ominous presence of the night-mare. It is likely that the early use of the English term nightmare was to describe intense SP (Orly & Haines, 2014). Culture-specific presentations of SP-related felt presence experiences typically involve a maleficent supernatural being, such as a witch or an evil spirit. Some examples found across cultures include the kanashibari demon in Japan (Fukuda et al, 1987; Arikawa, Templer & Brown, 1999); kokma in the West Indies (Ness, 1983); "old hag" in Newfoundland (Hufford, 1989); pandafeche in Italy (Jalal, Romanelli & Hinton, 2015); uqumangirniq among the Inuit of Baffin Island (Law & Kirmayer, 2005); and many others (for a comprehensive list of terms for SP experiences see Adler, 2011). Figure 3 illustrates a possible SP representation (Orly & Haines, 2014): a Japanese demon Yamachichi oppresses and inhales the breath of the sleeper. The first systematic cultural exploration of SP was done by Hufford (1989): he described a phenomenon specific to Newfoundland – the "old hag" witch attack. In his book Hufford discusses the tension in situating SP experiences somewhere between the 'cultural source hypothesis', wherein cultural interpretations and framing influence how an experience unfolds, and the 'experiential source hypothesis', where some invariant lived experiences, such as SP, may influence the development of a spiritual 9 interpretation and formation of cultural beliefs (Hufford 1989, 2005). Similar to this notion, McNamara and Bulkeley (2015) proposed an experiential hypothesis to describe how dreams and other dream-associated experiences, including visions and transcendental experiences, can be seen as a cornerstone and a source of religious belief (McNamara & Bulkeley, 2015). According to this view, a number of cultural, religious and paranormal beliefs are shaped primarily by direct experience and then framed within a particular tradition, which imbues them with existential and metaphysical meaning, a notion that is reminiscent of William James' grounding of mystical experience in the phenomenology of lived experience (James, 1985). The effect of framing such intense subjective experiences within a cultural tradition can have at least two kinds of potentially opposing effects. On the one hand, many cultures provide not only supernatural explanations of SP, but also remedies and protective rituals against it (some of which are described in the later part of this chapter), thus rooting the SP in a framework which allows for shared narrative and for practical interventions. On the other hand, intense and fearful SP, when interpreted as supernatural assault, has a potential for traumatizing the sleeper, thus creating a vicious circle of anxiety, aversion to sleep, facilitation of future SP episodes (Hinton et al, 2005; Sharpless et al, 2010), and increasing the level of distress via "cultural fear priming" (Ohayon et al, 1999; Jalal, Romanelli, & Hinton, 2015). For instance, the Incubus experience, when seen as part of the Christian tradition starting with the late Antique period, according to Gordon (2015), gained additional stigmatizing power, with a connotation of an illicit supernatural sexual experience. Not only were SP victims living through a waking nightmare of an encounter with a demonic assailant, they were also seen as responsible for having summoned it due to their own sinful predisposition/thoughts/impurities. It is important to note that while SP can include a range of experiences, such as positive experiences, neutral emotions, vestibulo-motor phenomena, out-of-body experiences, and others, most cultural interpretations of SP deal specifically with overlapping aspects of Intruder and Incubus. Figure 3. Takehara Shunsen. Yamachichi. Public domain image: https://commons.wikimedia.org/w/index.php?curid=2074508 Human sleep is typically divided into two kinds: REM sleep and NREM sleep. Healthy adults alternate between NREM and REM in cycles lasting about 90 minutes, for a total of 5-6 cycles over a night of sleep. Neural basis, associated conditions, and precipitating factors 10 While it is possible to experience dreaming in all stages of sleep, REM sleep is typically characterized by the most vivid, realistic, bizarre, and emotionally intense sleep mentation (Nielsen, 2000). Other vivid dream experiences, such as nightmares (Nielsen & Levin, 2007) and lucid dreams (LaBerge, Levitan & Dement, 1986), are also typically associated with REM sleep. Within the context of narcolepsy, sleep paralysis is a part of the diagnostic tetrad, alongside daytime sleepiness, cataplexy, and hypnagogic hallucinations (Thorpy, 2016). There is not sufficient data to assess whether there are significant differences in phenomenology between narcolepsy-associated SP and the isolated form. SP episodes are characterized by simultaneous presence of waking thought and of REM sleep psychophysiology (Mahowald & Schenck, 1991, 2005; Terzaghi et al., 2012), and the sleeper can often open her eyes and become relatively aware of her environment, while REM sleep-related spontaneous mentation - vivid dreaming - superimposes onto otherwise awake consciousness. This imagery may occur at sleep onset (hypnagogic) or upon awakening (hypnopompic). Other characteristics of REM sleep, such as airway occlusion and rapid shallow respiration (Gould et al., 1988) may contribute to the feeling of being suffocated or the perception of shortness of breath often reported by SP sufferers. Additionally, in one study obstructive sleep apnea was found to be a possible precipitating factor for ISP (Hsieh et al, 2010). Little research has been done on the sleep characteristics of SP sufferers. Some preliminary data suggests that the SP sleep profile may be similar to that of frequent nightmare sufferers (Nielsen et al, 2010), in that SP participants appear, paradoxically, to exhibit less REM sleep pressure, have more "skipped" REM sleep periods, and show no increase in eye movement density (as opposed to healthy controls) throughout the night (Solomonova, et al, 2012). SP participants also show higher delta power during sleep than non-SP controls (Marquis et al, 2015), which suggests alteration of processes of wake-NREM-REM regulation. Some of the vestibulo-motor characteristics, such as autoscopy, out-of-body experiences, and feelings of physical transformation, may stem from disturbances in right parietal regions (Jalal & Ramachandran, 2014): the mismatch between intended motor movement and inability to move may contribute to unusual physical sensations. SP may be experimentally elicited in laboratory settings, but only using an arduous protocol of repeated sleep interruption. For example, SP episodes were experimentally induced by letting participants sleep uninterrupted for the first NREM period, thus eliminating most of the slow-wave sleep pressure (a tendency of slow-wave NREM sleep to take precedence and occupy a large proportion of early night sleep), and then repeatedly awakening participants after 5 minutes of REM sleep have elapsed, thus augmenting REM sleep pressure and facilitating sleep-onset REM periods (SOREMPs). SOREMPs may be seen as a facilitating factor in a REM-wake state dissociation thought to characterize SP experiences. It should be noted, however, that even within such controlled settings and demanding protocols, rates of SP were relatively low: 6 episodes total in 16 participants who already had a tendency toward recurrent isolated SP (Takeuchi et al., 1992), and 8 episodes from 184 sleep interruptions in 13 SP sufferers (Takeuchi et al, 2002). These results suggest that incidence of SP at sleep onset may signify an individual's propensity to enter into REM sleep directly upon falling asleep. This further supports the idea that SP may result from alterations in wake-REM- NREM regulation patterns, resulting in state overlap. Little is known about the epidemiology of SP, but growing evidence points to a combination of genetic and experiential factors. The only study to date to examine genetic factors associated with SP has reported moderate heritability and that this effect was associated with factors known to contribute to disrupted sleep cycles (Denis et al., 2015). Sleep fragmentation and disruption in wake-NREM-REM regulation are an important factor facilitating SP occurrence, but it is uncertain whether all types of SP can be explained by a propensity for sleep fragmentation. Some ethnic groups seem to be more likely to experience SP than others. The Hmong population in Wisconsin, for instance, had a significantly higher incidence of SP than a non-Hmong cohort (Young et al, 2013), with as much as 31% of interviewed Hmong participant reporting at least weekly occurrence of SP episodes. Individuals of African descent also seem to have elevated rates of SP (Bell et al, 1984; Friedman & Paradis, 2002). Links between affective disorders, especially depression and anxiety, and SP have also been reported. A relationship have been found between SP and depression magnitude and anxiety (Szklo-Coxe, Young, Finn, & Mignot, 2007), social phobia and panic disorder (Paradis & Friedman, 2005; Otto et al, 2006; Sharpless et al, 2010) and social anxiety (Simard & Nielsen, 2005), especially with the sensation of being observed Associated conditions 11 (Solomonova et al, 2008). Changes in REM sleep regulation are often found in mood disorders, especially in depression (Arargun & Cartwright, 2003; Nofzinger et al, 1994) The relationship between trauma, especially post-traumatic stress disorder (PTSD), and SP has been noted by a number of researchers. McNally and Clancy found that there was a higher proportion of SP reports in participants with a history of childhood sexual abuse (McNally & Clancy, 2005a), and Abrams and colleagues (2008) reported that sexual abuse survivors report more distressing and more frequent SP incidence. In addition, higher rates of SP were found in Hmong population in relation to traumatic Vietnam War experiences (Young et al, 2013), as well as in Khmer (Hinton et al, 2005a) and Cambodian refugees (Hinton et al, 2005b). Similarly, Sharpless and Grom (2013) report that some cases of SP onset in adolescents begin after the loss of a family member. Considering that SP may be conceptualized as a nightmare spectrum experience, this relationship may represent the same dysfunction in the affect regulation network (Levin and Nielsen, 2007; Nielsen & Levin, 2007) as the one that has been proposed to be involved in nightmare production. PTSD-related sleep disturbances have been extensively documented (Spoormaker & Montgomery, 2008; Germain, Buysse & Nofzinger, 2008), including REM sleep dysregulation and increased nightmares (Melman et al, 2002; Germain, 2013), which in itself may contribute to altered REM sleep pressure, in turn facilitating occurrence of SP episodes. Since SP is often associated with intense, detailed, and troubling visions, a link between SP and psychiatric disorders has been hypothesized. Research, however, shows no consistent relationship between psychiatric conditions and SP, with the exceptions of PTSD, panic disorder, and social anxiety. In one study a number of links between SP and psychiatric conditions were found (Ohayon et al., 1999); these findings,were challenged, however, by an internet-based study (Solomonova et al., 2008), with a larger sample size, in which no strong links between psychopathology and SP were described. However, while isolated SP often presents itself in the absence of psychopathology, higher rates of hypnagogic and hypnopompic experiences (dream experiences occurring during the transition between sleep and wake: at sleep onset or upon awakenings, respectively), some of which may be associated with SP, are often found in psychosis (Plante and Winkelman, 2008). In their recent book, Sharpless and Doghramji (2015) list a number of plausible precipitating factors for SP occurrence in susceptible individuals. Sleep fragmentation and insufficient sleep are among the most obvious factors. REM sleep deprivation has been shown to increase REM sleep pressure contributing to REM rebound effect and intensified dreams at sleep onset (Nielsen et al, 2005). Poor sleep quality with frequent awakenings and disruptions may also facilitate REM-wake overlap, creating fruitful conditions for the occurrence of SOREMPs (Takeuchi et al, 1991, 2002; Spanos et al, 1995). Shift work, jet-lag, use of sleep disrupting medication, stress, anxiety – all these factors affect sleep and may facilitate a SP episode. Alcohol consumption was also reported to promote SP (Golzari & Ghabili, 2013; Munezawa et al, 2011), probably due to its effect on altering sleep architecture (Roehrs & Roth, 2001). Sleeping in a supine position also appears to enhance the risk of a SP episode (Sharpless et al, 2010). While undoubtedly felt presences are a hallmark of SP, especially of the intense and frightening episodes, presence experiences are not restricted to this parasomnia and are reported in a variety of conditions, thus possibly representing a more general and basic social imagery process (Nielsen 2007, Solomonova, Frantova & Nielsen, 2011). Arguably, the most salient and compelling felt presence occurs in the context of mystical and spiritual experiences. Otto (1958) introduced the idea of the numinous as a cornerstone of religious mystical experiences. Some of the recent work comes from anthropology: the ecstatic presence of God is manifested in the community of Evangelical Christians in the USA (Luhrmann, 2012). Other examples of felt presence have been documented in situations that are physically and emotionally straining or novel. Some examples of these experiences include high altitude climbing (Brugger, Regard, Landis, & Oelz, 1999); feeling of the presence of a baby in postpartum mothers (Nielsen & Paquette, 2007); presence of deceased relatives in the context of bereavement (Simon-Buller, Christopherson & Jones, 1989; Taylor, 2005; Keen, Murray & Payne, 2013); in extreme environments, such as solitary sailing (see also chapter by Suedfeld Neurocognitive considerations Precipitating factors A return to felt presence 12 Toward a cultural neurophenomenology of SP in this volume), surviving in remote and hostile environments (Suedfeld & Mocellin, 1987), and others. While in most cases felt presence is experienced spontaneously, in some cases it may be a product of sustained mental practices (as in prayer and some forms of meditation). One contemporary non-religious phenomenon is tulpamancy (Veissiere, 2016) – a long-term practice of conjuring up imaginary companions, that, over time, may be experienced as almost as real as other people. Additionally, being able to have a felt sense of others may be seen as a prerequisite for the development of subjectivity. Recent work in phenomenology and enactivism suggests that development of sense of self depends crucially on sensing others, as early as in utero (Gallagher, 2005; Ammaniti & Gallese, 2014), that the sense of one's own body depends on the sense of others (Maclaren, 2008) and that the self- other dynamic is a necessary condition for the sense of self (Zahavi, 2014). Evidence from dream research too suggests that dream processes are relational and intersubjective. The fact that dreams are most often about other people has been conceptualized as simulations of social interaction (Revonsuo, 2016) and as representations of individual attachment styles (McNamara et al, 2001). Additionally, dreams, similarly to waking, can be seen as a dynamic interaction between the "self"-related and "non-self" elements of dream content (everything extraneous to the dreamer). These non-self elements (non-human characters, dream environment, even dream objects can be seen as a "dream other" due to their inherent relational property (Solomonova et al 2015) and to the fact that dream environment in its entirety affectively motivates dreamer to engage with it. . SP has often been characterized as dissociative (Terzaghi et al., 2012) state, since it effectively combines characteristics of 'waking' consciousness (self-awareness, access to autobiographical memory, ability to open eyes and perceive the environment) with REM-sleep phenomena, specifically muscle atonia/paralysis and mentation/dreams. This notion of SP as dissociative has been at the heart of the previous neurobiological work on the link between dreaming and REM sleep. The relative deactivation of the dorsolateral prefrontal cortex characteristic of REM sleep (Hobson, Stickgold & Pace-Schott, 1988; Maquet, 2000) has been long hypothesized to be at the root of the loss of autobiographic memory and of the inability to appreciate the contents of the dream as "bizarre" or implausible in relation to reality. This has led to the hypothesis that in REM sleep dreaming one is effectively delusional and in a state of a transient psychosis (Hobson, 2004). In SP, similarly, there is often incomplete autobiographical access. This association between SP and REM sleep has also displaced the experience of SP from the psycho-spiritual domain of meaningful encounters with menacing/unreal/supernatural others, into a more reductionist account of uncontrollable and inescapable REM-initiated hallucinations. In contrast, an account of SP in the context of an oneiric phenomenology and in a 4EA perspective may allow for a more nuanced reading of these experiences. An emerging neurophenomenological framework of sleep challenges strict distinctions between wake, NREM, and REM sleep. Indeed, while SP is one of the examples of simultaneous presence of REM sleep and wake processes, it is not the only phenomenon that attests to the fluidity and interpermeability of states of consciousness. Lucid dreaming is another example of REM-wake co-occurrence (LaBerge, 1986); REM sleep behaviour disorder is characterized by preserved motor output during REM dreaming (Peever, Luppi & Montplaisir, 2014); somnambulism episodes combine NREM and wake physiology and phenomenology (Zadra et al, 2004); and a variety of dream-enacting behaviours, such as laughing, simple movement, crying and looking for a baby in bed, are prevalent in normal populations (Nielsen & Paquette, 2007; Nielsen, Svob & Kuiken, 2010). A more continuous view of mentation in sleep includes viewing SP as a form of oneiric experience: as a process of intensified mind-wandering (Fox et al, 2013), as a process of creativity (Hartmann & Kunzendorf, 2013), or as enactive imagination (Thompson, 2014), a process of sense-making in a rich, embodied and intersubjective world (Solomonova & Sha, 2016). In his discussion of lucid dreaming, Thompson (2014) proposes that in addition to seeing this state as a dissociative superimposition of two distinct states of consciousness, it may be simultaneously approached as an integrative state, thus allowing for an integration of two different yet related ways of self-experience. While SP sufferers feel awake and in their own bed, the realism of the experience and the quality of total immersion are completely overpowering to the dreamer, so that she is unable to appreciate the dreamlike quality or the unreality of the SP episode. The high prevalence of tactile and physical sensations 13 waking. But their presence seems so real, I would compare the experience to having them bizarre or terrifying event is happening all around me, and I am completely unable to respond or defend myself. Sometimes I know it's not real, somewhere in my mind, but it looks real, and it sounds real, and I'm terrified or revolted (or maybe just bemused), but I cannot wake myself up probably contributes to this effect. There are, however, numerous accounts of long-time SP experiencers that are characterized by a certain 'feel' for the experience as somewhere between real and unreal. SP-related experiences may have a very compelling and realistic quality, but they are usually lived differently from waking experiences, as a kind of a liminal state. Consider the following example: while the participant is experiencing intense emotion and is quite absorbed in the unfolding on the SP, he seems to have a kind of a dual awareness regarding the nature of his SP: "…Can't. Move. Not a muscle. Not an eyelash. It's often accompanied by hallucinations. So this to stop it." (30-year-old man, USA) Similarly, in another example the experiencer is also hesitant to ascribe any particular state to her experience: "… I might be answering wrong, because I see the beings in my dream-state immediately before accompanying me in the room". (48-year-old woman, USA) Grounding SP in its cultural context allows us to appreciate the variety of factors contributing to qualities of the lived experience, and it may not be possible to dissect the relative contribution of the multitude of neural, phenomenological and cultural narrative factors (Kirmayer, 2009). Importantly, in the current medical context, reducing SP to a dysfunction of REM psychophysiology may also have an important effect on reducing the potential for a deeper exploration of SP as a spiritual experience (Hufford, 2005). The cultural neurophenomenology of SP is a powerful tool for investigating SP from the 4EA cognition perspective. As neurophysiological, experiential accounts of SP show, the dreamer is in fact embodied – the oneiric scenario is dependent on the dreamer's state of consciousness (REM intrusion) and on the dreamer's physiological state (atonia, shallow rapid breathing). She is embedded in a physical (interprets ambiguous stimuli around her) and in a cultural world (these ambiguous stimuli take on a familiar shape/are infused with a deeper cultural and interpersonal signification). The sleeper is also extended into the world – the whole environment, both dreamt and real, is part of her ongoing experience; and her experience is enactive – there is a relational quality: she is not a passive observer of the oneiric drama unfolding before her eyes, but rather she is deeply engaged (Solomonova & Sha, 2016). In order to elucidate neurophenomenological qualities of SP in greater detail, future work may use microdynamic phenomenology/elicitation interviews, aimed at uncovering the fine-grained temporal and structural qualities of lived experience (Nielsen, this volume; Petitmengin, 2006; Petitmengin & Lachaux, 2013), in addition to neurophysiological data and deep awareness of the cultural, religious and spiritual context of the experiencer. While SP remains a relatively unknown phenomenon in much of Europe and America, a number of practical culture-specific practices have been developed to protect the sleeper from the negative influence of presumed supernatural forces. While some of these methods have deep roots in their respective metaphysical contexts, and therefore need to be grounded in existing religious and mystical practices, a number of practical and conceptually neutral recommendations have emerged, and seem beneficial for most SP sufferers, regardless of background. No established treatment for SP currently exists; its clinical management is instead often focused on treating comorbid problems. According to a review by Sharpless & Doghramji (2015), psychoanalysis, cognitive-behavioural therapy (CBT), hypnosis, and education in sleep hygiene have been investigated in relation to SP, but no empirical consensus on efficacy of such interventions is currently available. Based on the available evidence on SP and cognitive-behavioural approaches to treatment of sleep disorders, especially insomnia, the authors propose a manual for CBT-ISP. This is a promising first step toward finding a systematic method of dealing with SP. Sparse evidence for pharmacological interventions for SP also exists: in one study it was suggested that REM sleep-suppressing antidepressants may provide temporary relief (Plante and Winkelman, 2008), and treatment of narcolepsy may reduce SP frequency (Mamelak et al., 2004). Sleep paralysis practices: prevention, disruption, treatment and exploration 14 Antidepressants and anxiolytics were also used in severe cases (Hsieh et al., 2010). Terrillon and Marques- Bonham (2001) proposed that management of SP might benefit from administration of melatonin, which would help normalize the circadian rhythm. The cost of side effects associated with these treatments, however, may outweigh the benefit, and Shapless and Doghramji (2015) argue for a cautious approach, tailored to each individual situation. While methods for dealing with sleep paralysis have not been systematically explored by empirical psychology or cognitive science, the contemporary context of Internet-facilitated support groups and information sharing practices are changing the solitary and culture-bound nature of SP attacks. Furthermore, a number of methods have been anecdotally reported and documented online and in print, that see SP experiences as an opportunity rather than a nuisance, and promote exploration of one's own consciousness via SP-supported lucid dreaming or even contemplative approaches to SP (Hurd, 2010). One popular support group-mailing list is known as "Awareness during sleep paralysis" (ASP), and a reddit group on SP counts over 4000 users, sharing information on the phenomenology of their experiences and methods of overcoming them. Cultural and clinical practices associated with SP can be roughly separated into three kinds: 1) preventative practices, focused on avoiding SP-enabling circumstances; 2) disruptive practices, designed to stop SP in the middle of the experience; and 3) observational/explorative practices, aiming at observing SP and possibly transforming it into a positive event, such as a lucid dream or an out-of-body experience. Raising awareness of SP-associated phenomena itself may be one of the most important factors in reducing fear and distress before, during and after SP occurrence (Otto et al., 2006; Sharpless et al, 2010). Indeed, knowing that the experience is transient (will not last), benign (does not contain any real danger), and common (is shared with many individuals across the world) are powerful tools for psychological distancing and for facilitating an eventual observational, as opposed to fully immersive and fatalistic, attitude toward SP. Knowing about SP phenomenology and neurophysiology and having access to cultural grounding with available symbolic gestures helps prevent, disrupt and transform a negative experience into a tool for self-exploration. Figure 4 summarizes the intricate links between precipitating factors and effects of SP experiences in light of disruptive, and observational/transformational practices. While undoubtedly helpful, simply knowing the basis of SP may not be enough to alleviate terror and distress associated with the experiences, and disruption techniques are clearly warranted. A 25-year-old man from the United States reports: "This happens sometimes every night, sometimes only once every few weeks. Even though I 'know' what is happening, and that I am in no danger, it is always terrifying". The first study to systematically assess prevention strategies for SP by Sharpless and Grom (2014) has suggested that while no foolproof method for preventing SP is yet known, some strategies, such as avoiding sleeping on one's back (supine position), maintaining optimal sleep hygiene (avoiding stimulants, noise, irregular sleep patterns and anything that contributes to sleep fragmentation), and pre-sleep relaxation practices may help in preventing SP. A number of culture-specific preventative ritualistic measures to prevent SP exist. These include placing a variety of defensive objects in the room or in the bed before going to sleep, such as a variety of knives (Hufford, 1982, Law & Kirmayer, 2005); sprinkling salt (a common anti-witch remedy) (Roberts, 1998); putting a broom bottom-up (Paradis & Friedman, 2005) or a pile of sand at the bedroom door (Jalal, Romanelli & Hinton, 2015); and many others. Putting a Bible in the room (Hufford, 1982) and saying a protective prayer before bedtime are also thought of as effective deterrents. Other ritualistic actions, designed to deter, divert and chase away unwelcome supernatural visitors were also documented in a variety of contexts (Sharpless & Doghramji, 2015). While preventative measures, whether culturally embedded or aimed at increased awareness and promotion of sleep hygiene, may be effective in reducing the frequency of SP episodes, many methods for dealing with an ongoing SP experience also exist. Considering that most SP experiences are characterized by Methods for preventing sleep paralysis Techniques for disrupting sleep paralysis 15 Observational/transformational practices fear and other unpleasant sensations, it is not surprising that in one study the majority of participants reported having attempted to disrupt the ongoing SP experience. Moving the extremities and self-monitoring (raising awareness, promoting calm) may be helpful during the SP episode (Sharpless & Grom, 2014). Not all attempts or all strategies are equally successful, but it seems that attempting micro movements, instead of trying to get up or to scream, are most effective. Culture-bound rituals include saying a prayer (Hufford, 1982), making a sign of a cross with one's tongue (Davies, 2010), and asking someone to physically shake the oppressed sleeper (Law and Kirmayer, 2005). One may argue that ISP and lucid dreaming are polar opposites. However, they share the same underlying psychophysiology and seem to involve similar mechanisms: both are dependent upon REM sleep mechanisms; both are characterized by simultaneous presence of the dream state and by the feeling of being awake, including activation of higher order metacognitive functions indicative of some degree of waking thought processes (LaBerge, Levitan & Dement, 1986; Voss, Holzmann, Tuin, & Hobson, 2009; Dresler et al, 2012; Filevich et al, 2015); and in both cases muscle atonia is present. The crucial difference between the two states is the quality and the focus of awareness and metacognition: in lucid dreaming one is aware of the illusory nature of the dream scenario, whereas in SP the dreamer is often absorbed by the vision, not always fully realizing that it is dreamlike, and, in case of fearful SP, is too absorbed in the panicky state of perceived imminent danger. The link between SP and lucid dreaming has not been systematically investigated in empirical research, but two studies report a positive correlation between frequency of lucid dreaming and SP (Denis & Poerio, 2016; Solomonova, Nielsen & Stenstrom, 2009), suggesting that the REM-wake intertwined state, characterizing SP, may be a trait-like phenomenon predisposing individuals to SP on the one hand, and facilitating lucidity in REM sleep dreams on the other. Transforming SP into a positive experience, such as an OBE or a lucid dream, or utilizing SP experiences as a means of contemplative insight into one's own mind, may become a practice in itself, since not only techniques for disrupting and preventing SP exist in the contemporary digital culture, but also techniques for inducing SP, with the hope that the experience will function as a portal to a desirable altered state of consciousness (Hurd, 2010). The following two reports illustrate the transformative potential of SP: "I have woken up from dreaming and found I can't move or open my eyes. I get the feeling of lemonade bubbling in my body, especially my head. It is very frightening. But since I have been having OBE3s I now relax and go with the flow of sleep paralysis and sometimes I actually achieve an OBE" (40-year-old man, Australia) ignore and now I've been trying to communicate with the presence". (40-year-old man, USA) "At first I was very frightened until I found the ASP email group and found that I was not the only one being "visited" by this being during sleep paralysis. … When it first started happening it was more of an assault and I had to fight terribly to escape. But after years, I learned to 3 OBE = out-of-body experience 16 Further considerations and future directions Figure 4. Predisposing, precipitating factors and experience and outcome of sleep paralysis episodes. In terms of possible avenues for treatment, since SP can be conceptualized as a form of nightmare occurring in a mixed state of consciousness, nightmare treatment techniques could be useful in approaching SP. Currently, the most used and recommended technique for treating chronic nightmares is the Imagery Rehearsal Therapy (Krakow et al, 1995; Krakow & Zadra, 2006), which consists of "rehearsing" and transforming dysphoric oneiric imagery in a safe context. This method has been effective in treating PTSD- related nightmares (Krakow et al, 2001; Germain et al, 2004; Cook et al, 2010; Casement & Swanson, 2012), which seems particularly appropriate for intense and trauma-related SP experiences. Similarly, treatment of nightmares by lucid dreaming is a promising avenue (Zadra & Pihl, 1997; Spoormaker & Van Den Bout, 2006; LaBerge, 2009). Considering that neurophysiologically both states are characterized by an overlap between REM sleep and wakefulness, and that a number of folk approaches treating SP as a portal to lucid dreams already exist, mastering lucid dreaming could be an effective approach to transformation of an ongoing SP episode. Such a strategy may also be highly effective in de-stigmatizing and desensitising the experiencer, and especially in increasing her mastery and agency over her spontaneous oneiric experiences. Contemplative practices, such as meditation or pranayama (yogic breathing) may also be useful in dealing with recurring SP episodes. There is currently no empirical evidence for contemplative techniques and SP management, with the exception of a case study by Jalal (2016), but anecdotal evidence from practitioners as well as growing empirical literature linking contemplative practices with stress management, emotion regulation, and increased self-awareness, provide grounds for future research. Recent years have seen an important increase in empirical studies on the effects of meditation and meditation-based mindfulness interventions. There are documented benefits of contemplative practice in clinical populations including positive effects in mood disorders such as anxiety and depression (Hoffman et al, 2010; Goyal et al, 2014), social anxiety (Goldin & Gross, 2010), and PTSD (Kearney et al, 2013). At least four kinds of meditation are currently investigated in relation to mental health: focused attention, open monitoring (Lutz et al, 2008), self-transcendence (Travis & Shear, 2010) and loving kindness meditation (Hoffman, Grossman & Hinton, 2011). Different kinds of meditation practices may recruit different neural networks (Fox et al, 2016), and particular psychological and neuroplastic changes, associated with meditation practice, likely depend on the kind and duration of meditation experience (Lutz et al, 2015). These different kinds of contemplative practice may be helpful in targeting different kinds of recurrent SP experiences, promoting de-automatization (Kang, Gruber & Gray, 2013): deconstructing patterns of behaviour/reactivity. 17 Meditation may be effective in SP management as a way of cultivating a non-judgemental or 'non-sticky' observational attitude to arising imagery, sensations and emotions, and in letting the experience unfold. In addition, one important feature of most mindfulness-related practices is the focus on the experience of the body (Kerr et al, 2013), and some evidence suggests that meditation practice may improve awareness of one's own body states (Solomonova et al, 2016) and increase introspective accuracy for somatic experience (Fox et al, 2012). Breathing practices, such as pranayama, may be particularly effective in transforming SP as it is happening due to the fact that many SP episodes are characterised by a feeling of disordered/insufficient breathing. A recent study (Seppälä et al, 2014) reported that breathing exercises were effective in decreasing PTSD symptoms in war veterans. This implies that practicing techniques that improve awareness of body sensations may lower the reactivity to SP episodes, thus lowering the distressing quality of the experience, and increasing the potential for disrupting or transforming SP. The author was supported by Social Sciences and Humanities Research Council (SSHRC) of Canada and by a J.- A. DeSèves Sacre-Coeur Hospital Foundation doctoral scholarship. Thanks are due to Tore Nielsen, Philippe Stenstrom and Michelle Carr for numerous conversations on sleep paralysis and its interpretation, and to the members of the Dream and Nightmare Laboratory at the Center for Advanced Research in Sleep Medicine. Additional thanks are due to Don Donderi and Elena Frantova, as well as to all participants who have consented to share their experiences with us. Special thanks to Benjamen Samaha (artist) and to the anonymous SP sufferer for generously offering the narrative and the drawing of a sleep paralysis episode. Acknowledgements: 18 References: Abrams, M. P., Mulligan, A. D., Carleton, R. N., & Asmundson, G. J. G. (2008). Prevalence and correlates of sleep paralysis in adults reporting childhood sexual abuse. Journal of Anxiety Disorders, 22(8), 1535–1541. doi:10.1016/j.janxdis.2008.03.007 Adler, S. R. (2011). Sleep paralysis: Night-mares, nocebos, and the mind-body connection. Rutgers University Press. American Academy of Sleep Medicine. (2014). International classification of sleep disorders-third edition (ICSD-3). Darien, Illinois. American Psychiatric Association. (2013). Diagnostic and statistical manual of mental disorders (DSM-5). American Psychiatric Pub. Ammaniti, M., & Gallese, V. (2014). The birth of intersubjectivity: Psychodynamics, neurobiology, and the self. WW Norton & Company. Antelmi, E., Ferri, R., Iranzo, A., Arnulf, I., Dauvilliers, Y., Bhatia, K. P., … Plazzi, G. (2016). From state dissociation to status dissociatus. Sleep Medicine Reviews, 28, 1–13. doi:10.1016/j.smrv.2015.07.003 Agargun, M. Y., & Cartwright, R. (2003). REM sleep, dream variables and suicidality in depressed patients. Psychiatry Research, 119(1), 33-39. Arikawa, H., Templer, D. I., & Brown, R. (1999). The structure and correlates of kanashibari. The Journal of …. Retrieved from http://www.tandfonline.com/doi/abs/10.1080/00223989909599749 Belicki, K. (1992). Nightmare frequency versus nightmare distress: relations to psychopathology and cognitive style. Journal of abnormal psychology, 101(3), 592. Belicki, K., Gulko, N., Ruzycki, K., & Aristotle, J. (2003). Sixteen years of dreams following spousal bereavement. OMEGA-Journal of Death and Dying, 47(2), 93-106. Bell, C. C., Shakoor, B., Thompson, B., Dew, D., Hughley, E., Mays, R., & Shorter-Gooden, K. (1984). Prevalence of isolated sleep paralysis in black subjects. Journal of the National Medical Association, 76(5), 501. Blagrove, M., Farmer, L., & Williams, E. (2004). The relationship of nightmare frequency and nightmare distress to well-being. Journal of sleep research, 13(2), 129-136. Bloom, J. D., & Gelardin, R. D. (1983). Uqamairineq and Uqumanigianiq: Eskimo Sleep Paralysis. The Culture- Bound Syndromes, 117–122. doi:10.1007/978-94-009-5251-5_10 Brugger, P., Regard, M., & Landis, T. (1996). Unilaterally Felt" Presences": The Neuropsychiatry of One's Invisible Doppelganger. Cognitive and Behavioral Neurology, 9(2), 114-122. Brugger, P., Regard, M., & Landis, T. (1997). Illusory reduplication of one's own body: phenomenology and classification of autoscopic phenomena.Cognitive Neuropsychiatry, 2(1), 19-38. Brugger, P., Regard, M., Landis, T., & Oelz, O. (1999). Hallucinatory Experiences in Extreme-Altitude Climbers. Cognitive and Behavioral Neurology, 12(1), 67. Retrieved from http://journals.lww.com/cogbehavneurol/Fulltext/1999/01000/Hallucinatory_Experiences_in_Extre me_Altitude.8.aspx Buzzi, G. (2011). False awakenings in light of the dream protoconsciousness theory: a study in lucid dreamers. International Journal of Dream Research, 4(2), 110-116. Casement, M. D., & Swanson, L. M. (2012). A meta-analysis of imagery rehearsal for post-trauma nightmares: effects on nightmare frequency, sleep quality, and posttraumatic stress. Clinical psychology review, 32(6), 566-574. Cheyne, J. A. (2001). The ominous numinous. Sensed presence and 'other' hallucinations. Journal of Consciousness Studies, 8(5-6), 133-150. Cheyne JA (2002) Waterloo Unusual Sleep Experiences Questionnaire-VIIIa: Technical report. http://watarts.uwaterloo.ca/ *acheyne/spdoc/Techreport.pdf. Accessed 15 July 2016 Cheyne, J. A. (2005). Sleep paralysis episode frequency and number, types, and structure of associated hallucinations. Journal of Sleep Research, 14(3), 319–324. doi:10.1111/j.1365-2869.2005.00477.x Cheyne, J. A., & Girard, T. A. (2007). Paranoid delusions and threatening hallucinations: A prospective study of sleep paralysis experiences, 16(4), 959–974. doi:10.1016/j.concog.2007.01.002 Cheyne, J. A., Rueffer, S. D., & Newby-Clark, I. R. (1999). Hypnagogic and Hypnopompic Hallucinations during Sleep Paralysis: Neurological and Cultural Construction of the Night-Mare. Consciousness and Cognition, 8(3), 319–337. doi:10.1006/ccog.1999.0404 Cheyne, J. A., & Pennycook, G. (2013). Sleep Paralysis Postepisode Distress Modeling Potential Effects of Episode Characteristics, General Psychological Distress, Beliefs, and Cognitive Style. Clinical 19 Psychological Science, 2167702612466656. Clark, A., & Chalmers, D. (1998). The extended mind. analysis, 58(1), 7-19. Colombetti, G. (2013). The feeling body: Affective science meets the enactive mind. MIT Press. Davies, O. (2010). The Nightmare Experience, Sleep Paralysis, and Witchcraft Accusations. Folklore, 114(2), 181–203. doi:10.1080/0015587032000104211 Cook, J. M., Harb, G. C., Gehrman, P. R., Cary, M. S., Gamble, G. M., Forbes, D., & Ross, R. J. (2010). Imagery rehearsal for posttraumatic nightmares: a randomized controlled trial. Journal of traumatic stress, 23(5), 553-563. Denis, D., French, C. C., Rowe, R., Zavos, H. M. S., Nolan, P. M., Parsons, M. J., & Gregory, A. M. (2015). A twin and molecular genetics study of sleep paralysis and associated factors. Journal of Sleep Research, n/a. doi:10.1111/jsr.12282 Denis, D., & Poerio, G. L. (2016). Terror and bliss? Commonalities and distinctions between sleep paralysis, lucid dreaming, and their associations with waking life experiences. Journal of Sleep Research. Dresler, M., Wehrle, R., Spoormaker, V. I., Koch, S. P., Holsboer, F., Steiger, A., ... & Czisch, M. (2012). Neural correlates of dream lucidity obtained from contrasting lucid versus non-lucid REM sleep: a combined EEG/fMRI case study. Sleep, 35(7), 1017-1020. Farhall, J., Greenwood, K. M., & Jackson, H. J. (2007). Coping with hallucinated voices in schizophrenia: a review of self-initiated strategies and therapeutic interventions. Clinical Psychology Review, 27(4), 476-493. Filevich, Elisa, Martin Dresler, Timothy R. Brick, and Simone Kühn. "Metacognitive mechanisms underlying lucid dreaming." The Journal of Neuroscience 35, no. 3 (2015): 1082-1088. Fox, K. C., Zakarauskas, P., Dixon, M., Ellamil, M., Thompson, E., & Christoff, K. (2012). Meditation experience predicts introspective accuracy. PloS one,7(9), e45370. Fox, K. C., Nijeboer, S., Solomonova, E., Domhoff, G. W., & Christoff, K. (2013). Dreaming as mind wandering: evidence from functional neuroimaging and first-person content reports. Frontiers in human neuroscience, 7, 412. Fox, K. C., Dixon, M. L., Nijeboer, S., Girn, M., Floman, J. L., Lifshitz, M., ... & Christoff, K. (2016). Functional neuroanatomy of meditation: A review and meta-analysis of 78 functional neuroimaging investigations. Neuroscience & Biobehavioral Reviews, 65, 208-228. Friedman, S., & Paradis, C. (2002). Panic disorder in African-Americans: symptomatology and isolated sleep paralysis. Culture, medicine and psychiatry, 26(2), 179-198. Fukuda, K. (1993). One explanatory basis for the discrepancy of reported prevalences of sleep paralysis among healthy respondents. Perceptual and Motor Skills. Retrieved from http://www.amsciepub.com/doi/pdf/10.2466/pms.1993.77.3.803 Fukuda, K., Miyasita, A., Inugami, M., & Ishihara, K. (1987). High prevalence of isolated sleep paralysis: Kanashibari phenomenon in Japan. SLEEP. (PAGES) Fukuda, K., Ogilvie, R. D., & Takeuchi, T. (2000). Recognition of sleep paralysis among normal adults in Canada and in Japan. Psychiatry and clinical neurosciences, 54(3), 292-293. Gallagher, S. (2005). How the body shapes the mind (pp. 173-178). Oxford: Clarendon Press. Gordon, S. (2015). Medical Condition, Demon or Undead Corpse? Sleep Paralysis and the Nightmare in Medieval Europe. Social History of Medicine, 28(3), NaN–NaN. doi:10.1093/shm/hkv005 Garfield, P. (1996). Dreams and bereavement. In Trauma and Dreams, Barrett, D. editor. Pp.186-211 Germain, A., Krakow, B., Faucher, B., Zadra, A., Nielsen, T., Hollifield, M., ... & Koss, M. (2004). Increased Mastery Elements Associated With Imagery Rehearsal Treatment for Nightmares in Sexual Assault Survivors With PTSD. Dreaming, 14(4), 195. Germain, A., Buysse, D. J., & Nofzinger, E. (2008). Sleep-specific mechanisms underlying posttraumatic stress disorder: integrative review and neurobiological hypotheses. Sleep medicine reviews, 12(3), 185-195. Germain, A. (2013). Sleep disturbances as the hallmark of PTSD: where are we now?. American Journal of Psychiatry, 170(4), 372-382. Goldin, P. R., & Gross, J. J. (2010). Effects of mindfulness-based stress reduction (MBSR) on emotion regulation in social anxiety disorder. Emotion,10(1), 83. Golzari, S. E., & Ghabili, K. (2013). Alcohol-mediated sleep paralysis: the earliest known description. Sleep Medicine, 14(3), 298. Gould, G. A., Gugger, M., Molloy, J., Tsara, V. M. S. C., Shapiro, C. M., & Douglas, N. J. (1988). Breathing pattern and eye movement density during REM sleep in humans. Am Rev Respir Dis, 138(4), 874-877. 20 Green, C. E., & McCreery, C. (1994). Lucid dreaming: the paradox of consciousness during sleep. Psychology Press. Hartmann, E., & Kunzendorf, R. (2013). Thymophor in Dreams, Poetry, Art and Memory: Emotion Translated into Imagery as a Basic Element of Human Creativity. Imagination, Cognition and Personality, 33(1), 165-191. Hinton, D. E., Pich, V., Chhean, D., & Pollack, M. H. (2005a). 'The ghost pushes you down': sleep paralysis-type panic attacks in a Khmer refugee population. Transcultural Psychiatry, 42(1), 46–77. Hinton, D. E., Pich, V., Chhean, D., Pollack, M. H., & McNally, R. J. (2005b). Sleep paralysis among Cambodian refugees: association with PTSD diagnosis and severity. Depression and Anxiety, 22(2), 47-51. Hishikawa, Y., & Kaneko, Z. (1965). Electroencephalographic study on narcolepsy. Electroencephalography and Clinical Neurophysiology, 18(3), 249-259. Hobson, J. A., Stickgold, R., & Pace-Schott, E. F. (1998). The neuropsychology of REM sleep dreaming. Neuroreport, 9(3), R1-R14. Hobson J. A., Pace-Schott E. F. & Stickgold R. (2000) Dreaming and the brain: Toward a cognitive neuroscience of conscious states. Behavioral and Brain Science 23(6): 793–842; discussion 904–1121 Hobson, A. (2004). A model for madness?. Nature, 430(6995), 21-21. Hofmann, S. G., Sawyer, A. T., Witt, A. A., & Oh, D. (2010). The effect of mindfulness-based therapy on anxiety and depression: A meta-analytic review. Journal of consulting and clinical psychology, 78(2), 169. Hofmann, S. G., Grossman, P., & Hinton, D. E. (2011). Loving-kindness and compassion meditation: Potential for psychological interventions. Clinical psychology review, 31(7), 1126-1132. Howell, M. J. (2012). Parasomnias: an updated review. Neurotherapeutics : The Journal of the American Society for Experimental NeuroTherapeutics, 9(4), 753–775. doi:10.1007/s13311-012-0143-8 Hsieh, S.-W., Lai, C.-L., Liu, C.-K., Lan, S.-H., & Hsu, C.-Y. (2010). Isolated sleep paralysis linked to impaired nocturnal sleep quality and health-related quality of life in Chinese-Taiwanese patients with obstructive sleep apnea. Quality of Life Research : An International Journal of Quality of Life Aspects of Treatment, Care and Rehabilitation, 19(9), 1265–1272. doi:10.1007/s11136-010-9695-4 Hufford, D. (1989). The terror that comes in the night: An experience-centered study of supernatural assault traditions (Vol. 7). University of Pennsylvania Press. Hufford, D. J. (2005). Sleep paralysis as spiritual experience. Transcultural Psychiatry, 42(1), 11–45. Hurd, R. (2010). Sleep Paralysis: A Guide to Hypnagogic Visions and Visitors of the Night. Hyena Press. Goyal, M., Singh, S., Sibinga, E. M., Gould, N. F., Rowland-Seymour, A., Sharma, R., ... & Ranasinghe, P. D. (2014). Meditation programs for psychological stress and well-being: a systematic review and meta-analysis. JAMA internal medicine, 174(3), 357-368. Jalal, B., & Ramachandran, V. S. (2014). Sleep paralysis and 'the bedroom intruder': the role of the right superior parietal, phantom pain and body image projection. Medical Hypotheses, 83(6), 755–757. doi:10.1016/j.mehy.2014.10.002 Jalal, B., Simons-Rudolph, J., Jalal, B., & Hinton, D. E. (2014). Explanations of sleep paralysis among Egyptian college students and the general population in Egypt and Denmark. Transcultural Psychiatry, 51(2), 158–175. doi:10.1177/1363461513503378 Jalal, B., Romanelli, A., & Hinton, D. E. (2015). Cultural Explanations of Sleep Paralysis in Italy: The Pandafeche Attack and Associated Supernatural Beliefs. Culture, Medicine, and Psychiatry, 1–14. doi:10.1007/s11013-015-9442-y Jalal, B. (2016). How to make the ghosts in my bedroom disappear? Focused-attention meditation combined with muscle relaxation (MR therapy)-a direct treatment intervention for sleep paralysis. Frontiers in psychology,7. James, W. (1985). The varieties of religious experience (Vol. 15). Harvard University Press. Jiménez-Genchi, A., Ávila-Rodríguez, V. M., Sánchez-Rojas, F., Vargas Terrez, B. E., & Nenclares-Portocarrero, A. (2009). Sleep paralysis in adolescents: the 'a dead body climbed on top of me' phenomenon in Mexico.Psychiatry and clinical neurosciences, 63(4), 546-549. Kang, Y., Gruber, J., & Gray, J. R. (2013). Mindfulness and de-automatization. Emotion review, 5(2), 192-201. Kearney, D. J., Malte, C. A., McManus, C., Martinez, M. E., Felleman, B., & Simpson, T. L. (2013). Loving-kindness meditation for posttraumatic stress disorder: A pilot study. Journal of Traumatic Stress, 26(4), 426-434. Keen, C., Murray, C. D., & Payne, S. (2013). A qualitative exploration of sensing the presence of the deceased following bereavement. Mortality,18(4), 339-357. Kerr, C. E., Sacchet, M. D., Lazar, S. W., Moore, C. I., & Jones, S. R. (2013). Mindfulness starts with the body: 21 somatosensory attention and top-down modulation of cortical alpha rhythms in mindfulness meditation. Frontiers in human neuroscience, 7, 12. Kirmayer, L. J. (2009). Nightmares, neurophenomenology and the cultural logic of trauma. Culture, medicine and psychiatry, 33(2), 323-331. Krakow, B., Kellner, R., Pathak, D., & Lambert, L. (1995). Imagery rehearsal treatment for chronic nightmares. Behaviour Research and Therapy, 33(7), 837-843. Krakow, B., & Zadra, A. (2006). Clinical management of chronic nightmares: imagery rehearsal therapy. Behavioral sleep medicine, 4(1), 45-70. LaBerge, S., Levitan, L., & Dement, W. C. (1986). Lucid dreaming: Physiological correlates of consciousness during REM sleep. Journal of Mind and Behavior. LaBerge, S. (2009). Lucid dreaming: A concise guide to awakening in your dreams and in your life. ReadHowYouWant. com. Law, S., & Kirmayer, L. J. (2005). Inuit interpretations of sleep paralysis.Transcultural psychiatry, 42(1), 93- 112. Levin, R., & Fireman, G. (2002). Nightmare prevalence, nightmare distress, and self-reported psychological disturbance. SLEEP-NEW YORK-, 25(2), 205-212. Levin, R., & Nielsen, T. A. (2007). Disturbed dreaming, posttraumatic stress disorder, and affect distress: a review and neurocognitive model. Psychological bulletin, 133(3), 482. Liddon, S. C. (1967). Sleep paralysis and hypnagogic hallucinations: Their relationship to the nightmare. Archives of General Psychiatry, 17(1), 88. Liskova, M., Janeckova, D., Kluzova Kracmarova, L., Mlada, K., & Buskova, J. (2016). The occurrence and predictive factors of sleep paralysis in university students. Neuropsychiatr Dis Treat, 12, 2957-2962. doi:10.2147/NDT.S115629 Luhrmann, T. M. (2012). When God talks back: Understanding the American evangelical relationship with God. Vintage. Lutz, A., Slagter, H. A., Dunne, J. D., & Davidson, R. J. (2008). Attention regulation and monitoring in meditation. Trends in cognitive sciences, 12(4), 163-169. Lutz, A., Jha, A. P., Dunne, J. D., & Saron, C. D. (2015). Investigating the phenomenological matrix of mindfulness-related practices from a neurocognitive perspective. American Psychologist, 70(7), 632. Mamelak, M., Black, J., Montplaisir, J., & Ristanovic, R. (2004). A pilot study on the effects of sodium oxybate on sleep architecture and daytime alertness in narcolepsy. Sleep, 27, 1327-1336. Mahowald, M. W., & Schenck, C. H. (1991). Status dissociatus: A perspective on states of being. SLEEP. Retrieved from http://psycnet.apa.org/psycinfo/1991-30903-001 Mahowald, M. W., & Schenck, C. H. (2005). Insights from studying human sleep disorders. Nature, 437(7063), 1279-1285. Maclaren, K. (2008). Embodied perceptions of others as a condition of selfhood?. Journal of Consciousness Studies, 15(8), 63-93. Maquet, P. (2000). Functional neuroimaging of normal human sleep by positron emission tomography. Journal of sleep research, 9(3), 207-232. McNally, R. (2011). Explaining 'Memories' of Space Alien Abduction and Past Lives: An Experimental Psychopathology Approach. Journal of Experimental Psychopathology. doi:10.5127/jep.017811 McNally, R. J., & Clancy, S. A. (2005a). Sleep paralysis in adults reporting repressed, recovered, or continuous memories of childhood sexual abuse. Journal of Anxiety Disorders, 19(5), 595–602. doi:10.1016/j.janxdis.2004.05.003 McNally, R. J., & Clancy, S. A. (2005b). Sleep paralysis, sexual abuse, and space alien abduction. Transcultural Psychiatry, 42(1), 113–122. McNamara, P., Andresen, J., Clark, J., Zborowski, M., & Duffy, C. A. (2001). Impact of attachment styles on dream recall and dream content: a test of the attachment hypothesis of REM sleep. Journal of sleep research, 10(2), 117-127. McNamara, P., & Bulkeley, K. (2015). Dreams as a source of supernatural agent concepts. Frontiers in Psychology, 6, 283. doi:10.3389/fpsyg.2015.00283 Mellman, T. A., Bustamante, V., Fins, A. I., Pigeon, W. R., & Nolan, B. (2002). REM sleep and the early development of posttraumatic stress disorder.American Journal of Psychiatry, 159(10), 1696-1701. Menary, R. (2010). The extended mind. Cambridge, MA: The MIT Press. Morris D. (2010) Empirical and phenomenological studies of embodied cognition. In: Gallagher S. & 22 Schmicking D. (eds.) Handbook of phenomenology and cognitive science. Springer, New York: 235–252. Munezawa, T., Kaneita, Y., Osaki, Y., Kanda, H., Ohtsu, T., Suzuki, H., . . . Ohida, T. (2011). Nightmare and sleep paralysis among Japanese adolescents: a nationwide representative survey. Sleep Med, 12(1), 56-64. doi: 10.1016/j.sleep.2010.04.015 Ness, R. C. (1983). The Old Hag Phenomenon as Sleep Paralysis: A Biocultural Interpretation. The Culture- Bound Syndromes, 123–145. doi:10.1007/978-94-009-5251-5_11 Nielsen, T. A. (2000). A review of mentation in REM and NREM sleep: "covert" REM sleep as a possible reconciliation of two opposing models. Behavioral and Brain Sciences, 23(06), 851-866. Nielsen, T., Stenstrom, P., Takeuchi, T., Saucier, S., Lara-Carrasco, J., Solomonova, E., & Martel, E. (2005). Partial REM-sleep deprivation increases the dream-like quality of mentation from REM sleep and sleep onset. Sleep 28(9), 1083. Nielsen, T. (2007). Felt presence: paranoid delusion or hallucinatory social imagery? Consciousness and Cognition, 16(4), 975. doi:10.1016/j.concog.2007.02.002 Nielsen, T., & Levin, R. (2007). Nightmares: a new neurocognitive model. Sleep Medicine Reviews, 11(4), 295- 310. Nielsen, T., & Paquette, T. (2007). Dream-associated behaviors affecting pregnant and postpartum women. Sleep, 30(9), 1162. Nielsen, T., Svob, C., & Kuiken, D. (2009). Dream-enacting behaviors in a normal population. Sleep, 32(12), 1629-36. Nielsen, T. A., Paquette, T., Solomonova, E., Lara-Carrasco, J., Popova, A., & Levrier, K. (2010). REM sleep characteristics of nightmare sufferers before and after REM sleep deprivation. Sleep medicine, 11(2), 172-179. Noë, A. (2004). Action in perception. Cambridge, MA: The MIT press. Nofzinger, E. A., Schwartz, R. M., Reynolds, C. F., Thase, M. E., Jennings, J. R., Frank, E., ... & Kupfer, D. J. (1994). Affect intensity and phasic REM sleep in depressed men before and after treatment with cognitive- behavioral therapy. Journal of Consulting and Clinical Psychology, 62(1), 83. O'Hanlon, J., Murphy, M., & Di Blasi, Z. (2011). Experiences of sleep paralysis in a sample of Irish university students. Irish Journal of Medical Science, 180(4), 917–919. doi:10.1007/s11845-011-0732-2 Ohayon, M. M., Zulley, J., Guilleminault, C., & Smirne, S. (1999). Prevalence and pathologic associations of sleep paralysis in the general population. Neurology, 52(6), 1194–1200. doi:10.1212/WNL.52.6.1194 Olry, R., & Haines, D. E. (2014). Kanashibari : a ghost's business. Journal of the History of the Neurosciences, 23(2), 192–197. doi:10.1080/0964704X.2013.862132 Otto, R. (1958). The idea of the holy (Vol. 14). Oxford University Press. Otto, M. W., Simon, N. M., Powers, M., Hinton, D., Zalta, A. K., & Pollack, M. H. (2006). Rates of isolated sleep paralysis in outpatients with anxiety disorders. Journal of Anxiety Disorders, 20(5), 687–693. doi:10.1016/j.janxdis.2005.07.002 Paradis, C. M., & Friedman, S. (2005). Sleep paralysis in African Americans with panic disorder. Transcultural psychiatry, 42(1), 123-134. Peever, J., Luppi, P. H., & Montplaisir, J. (2014). Breakdown in REM sleep circuitry underlies REM sleep behavior disorder. Trends in neurosciences,37(5), 279-288. Petitmengin, C. (2006). Describing one's subjective experience in the second person: An interview method for the science of consciousness.Phenomenology and the Cognitive Sciences, 5(3-4), 229-269. Petitmengin, C., & Lachaux, J. P. (2013). Microcognitive science: bridging experiential and neuronal microdynamics. Frontiers in human neuroscience,7, 617. Pessoa, L. (2013). The cognitive-emotional brain: From interactions to integration. MIT press. Plante, D. T., & Winkelman, J. W. (2008). Parasomnias: Psychiatric Considerations. Sleep Medicine Clinics, 3(2), 217–229. doi:10.1016/j.jsmc.2008.01.005 Portocarrero A (2009): Sleep paralysis in adolescents: The "a dead body climbed on top of me" phenomenon in Mexico. Psychiatry and Clinical Neurosciences 63: 546-549. Protevi, J. (2012). One More 'Next Step': Deleuze and Brain, Body and Affect in Contemporary Cognitive Science. Revisiting Normativity with Deleuze, 25. Rechtschaffen A. (1978) The single-mindedness and isolation of dreams. Sleep 1: 97–109 Revonsuo, A. (2006). Inner presence: Consciousness as a biological phenomenon. Mit Press. Revonsuo, A., Tuominen, J., Valli, K., Dresler, M., Metzinger, T. & Windt, J.M. "The Avatars in the Machine." Open MIND. Frankfurt a. M., GER: MIND Group (2015). 23 Roehrs, T., & Roth, T. (2001). Sleep, sleepiness, sleep disorders and alcohol use and abuse. Sleep medicine reviews, 5(4), 287-297. Robert, G., & Zadra, A. (2014). Thematic and content analysis of idiopathic nightmares and bad dreams. Sleep, 37(2), 409-417 Roberts, K. (1998). Contemporary cauchemar: Experience, belief, prevention. Louisiana Folklore Miscellany, 13, 15-26. Seppälä, E. M., Nitschke, J. B., Tudorascu, D. L., Hayes, A., Goldstein, M. R., Nguyen, D. T., ... & Davidson, R. J. (2014). Breathing-based meditation decreases posttraumatic stress disorder symptoms in US Military veterans: A randomized controlled longitudinal study. Journal of traumatic stress, 27(4), 397-405. Sharpless, B. A., McCarthy, K. S., Chambless, D. L., Milrod, B. L., Khalsa, S.-R., & Barber, J. P. (2010)., 66(12), 1292–1306. doi:10.1002/jclp.20724 Sharpless, B. A., & Barber, J. P. (2011). Lifetime prevalence rates of sleep paralysis: a systematic review. Sleep Medicine Reviews, 15(5), 311–315. doi:10.1016/j.smrv.2011.01.007 Sharpless, B. A., & Grom, J. L. (2014). Isolated Sleep Paralysis: Fear, Prevention, and Disruption. Behavioral Sleep Medicine, 1–6. doi:10.1080/15402002.2014.963583 Sharpless, B., & Doghramji, K. (2015). Sleep paralysis: Historical, psychological, and medical perspectives. Oxford University Press. Simard, V., & Nielsen, T. A. (2005). Sleep paralysis-associated sensed presence as a possible manifestation of social anxiety. Dreaming, 15(4), 245. Simon-Buller, S., Christopherson, V. A., & Jones, R. A. (1989). Correlates of sensing the presence of a deceased spouse. OMEGA-Journal of Death and Dying, 19(1), 21-30. Solomonova, E., Nielsen, T., Stenstrom, P., Simard, V., Frantova, E., & Donderi, D. (2008). Sensed presence as a correlate of sleep paralysis distress, social anxiety and waking state social imagery. Consciousness and Cognition, 17(1), 49–63. doi:10.1016/j.concog.2007.04.007 Solomonova, E., Nielsen, T., & Stenstrom, P. (2009). Lucid dreaming is associated with sleep paralysis but not nightmares. Third Annual Conference of the Canadian Sleep Society. April 26-28, Toronto, Canada Solomonova, E., Frantova, E., & Nielsen, T. (2010). Felt presence: the uncanny encounters with the numinous Other. AI & SOCIETY, 26(2), 171–178. doi:10.1007/s00146-010-0299-x Solomonova, E., Nielsen, T., Takeuchi, T. Bezinger, E. & Carr, M. (2012). The relationship between REM sleep eye movement density and dream content: differences between idiopathic sleep paralysis sufferers and healthy controls. International Association for the Study of Dreams annual meeting. June 24, 2012. Berkeley, CA. Solomonova, E., Stenstrom, P., Paquette, T., & Nielsen, T. (2015). Different temporal patterns of memory incorporations into dreams for laboratory and virtual reality experiences: relation to dreamed locus of control. International journal of dream research, 8(1), 10-26. Solomonova, E. & Sha, X.W. (2016). Exploring the depth of dream experience: The enactive framework and methods for neurophenomenological research. Constructivist Foundations. Special Issue: Exploring the Diversity within Enactivism and Neurophenomenology. Tom Froese, Alexander Riegler and Sebastian Voros, editors. 11(2), 407-416. Solomonova, E., Dubé, S., Blanchette-Carrière, C., Samson, A., Picard-Deland, C. & Nielsen, T. (2016). Effect of Vipassana meditation on sleep-dependent procedural memory consolidation and postural balance following a daytime nap: Making implicit memories explicit. Mind and Life Summer Research Institute, June 2016, Garrison, NY, USA. Spanos, N. P., McNulty, S. A., DuBreuil, S. C., Pires, M., & Burgess, M. F. (1995). The frequency and correlates of sleep paralysis in a university sample. Journal of Research in Personality, 29(3), 285-305. Spoormaker, V. I., & Van Den Bout, J. (2006). Lucid dreaming treatment for nightmares: a pilot study. Psychotherapy and psychosomatics, 75(6), 389-394. Spoormaker, V. I., & Montgomery, P. (2008). Disturbed sleep in post-traumatic stress disorder: secondary symptom or core feature?. Sleep medicine reviews, 12(3), 169-184. Stewart, J. R., Gapenne, O., & Di Paolo, E. A. (2010). Enaction: Toward a new paradigm for cognitive science. MIT Press. Suedfeld, P., & Mocellin, J. S. (1987). The" Sensed Presence" in Unusual Environments Peter Suedfeld. Environment and Behavior, 19(1), 33-52. Szelenberger, W., Niemcewicz, S., & Dabrowska, A. J. (2005). Sleepwalking and night terrors: psychopathological and psychophysiological correlates. International Review of Psychiatry, 17(4), 263- 24 270. Szklo-Coxe, M., Young, T., Finn, L., & Mignot, E. (2007). Depression: relationships to sleep paralysis and other sleep disturbances in a community sample. Journal of sleep research, 16(3), 297-312. Takeuchi, T., Miyasita, A., Sasaki, Y., Inugami, M.. (1992). Isolated sleep paralysis elicited by sleep interruption. SLEEP. (PAGES) Takeuchi, T., Fukuda, K., Sasaki, Y., Inugami, M., & Murphy, T. I. (2002). Factors related to the occurrence of isolated sleep paralysis elicited during a multi-phasic sleep-wake schedule. Sleep, 25(1), 89-96. Taylor, S. F. (2005). Between the idea and the reality: A study of the counselling experiences of bereaved people who sense the presence of the deceased. Counselling and Psychotherapy Research, 5(1), 53-61. Terrillon, J. C., & Marques-Bonham, S. (2001). Does recurrent isolated sleep paralysis involve more than cognitive neurosciences. Journal of Scientific Exploration, 15(1), 97-123. Terzaghi, M., Ratti, P. L., Manni, F., & Manni, R. (2012). Sleep paralysis in narcolepsy: more than just a motor dissociative phenomenon? Neurological Sciences : Official Journal of the Italian Neurological Society and of the Italian Society of Clinical Neurophysiology, 33(1), 169–172. doi:10.1007/s10072-011-0644-y Thompson, E. (2005). Sensorimotor subjectivity and the enactive approach to experience. Phenomenology and the cognitive sciences, 4(4), 407-427. Thompson, E. (2014). Waking, Dreaming, Being: Self and Consciousness in Neuroscience, Meditation, and Philosophy. New York: Columbia University Press. Thompson E. (2015a) Steps toward a neurophenomenology of conscious sleep. A Reply to Jennifer M. Windt. In: Metzinger T. & Windt J. M. (eds.) Open MIND. MIND Group, Frankfurt am Main: 37(R). http://open- mind.net/papers/steps-toward-a-neurophenomenology-of-consciousness-in-sleep-a-reply-to-jennifer- m-windt Thompson E. (2015b) Dreamless sleep, the embodied mind, and consciousness: The Relevance of a classical Indian debate to cognitive science. In: Metzinger T. & Windt J. M. (eds.) Open MIND. MIND Group, Frankfurt am Main: 37(T). http://open-mind.net/papers/dreamless-sleep-the-embodied-mind-and- consciousness-the-relevance-of-a-classical-indian-debate-to-cognitive-science Thorpy, M. J. (2016). Diagnostic Criteria and Delay in Diagnosis of Narcolepsy. In Narcolepsy (pp. 45-49). Springer International Publishing. Travis, F., & Shear, J. (2010). Focused attention, open monitoring and automatic self-transcending: categories to organize meditations from Vedic, Buddhist and Chinese traditions. Consciousness and cognition, 19(4), 1110-1118. Varela F. J., Thompson E. & Rosch E. (1991) The embodied mind: Cognitive science and human experience. MIT Press, Cambridge MA. Veissière, S. (2016). Varieties of Tulpa Experiences: The Hypnotic Nature of Human Sociality, Personhood, and Interphenomenality. In Raz and Lifshitz, eds. Hypnosis and Meditation Towards an Integrative Science of Conscious Planes. Oxford University Press. Wheeler, M. (2005). Reconstructing the cognitive world: The next step. MIT press. Wing, Y.-K., Chiu, H., Leung, T., & Ng, J. (1999). Sleep paralysis in the elderly. Journal of Sleep Research, 8(2), 151–155. doi:10.1046/j.1365-2869.1999.00143.x Wing, Y.-K., Lee, S. T., & Chen, C.-N. (1994). Sleep paralysis in Chinese: Ghost oppression phenomenon in Hong Kong. Sleep. Zadra, A. L., & Pihl, R. O. (1997). Lucid dreaming as a treatment for recurrent nightmares. Psychotherapy and Psychosomatics, 66(1), 50-55. Zadra, A., Pilon, M., Joncas, S., Rompré, S., & Montplaisir, J. (2004). Analysis of postarousal EEG activity during somnambulistic episodes.Journal of sleep research, 13(3), 279-284. Zahavi, D. (2014). Self and other: Exploring subjectivity, empathy, and shame. OUP Oxford. 25
1504.06290
1
1504
2015-04-23T18:36:01
Entrainment in up and down states of neural populations: non-smooth and stochastic models
[ "q-bio.NC" ]
We study the impact of noise on a neural population rate model of up and down states. Up and down states are typically observed in neuronal networks as a slow oscillation, where the population switches between high and low firing rates (Sanchez-Vivez and McCormick, 2000). A neural population model with spike rate adaptation is used to model such slow oscillations, and the timescale of adaptation determines the oscillation period. Furthermore, the period depends non-monotonically on the background tonic input driving the population, having long periods for very weak and very strong stimuli. Using both linearization and fast-slow timescale separation methods, we can compute the phase sensitivity function of the slow oscillation. We find that the phase response is most strongly impacted by perturbations to the adaptation variable. Phase sensitivity functions can then be utilized to quantify the impact of noise on oscillating populations. Noise alters the period of oscillations by speeding up the rate of transition between the up and down states. When common noise is presented to two distinct populations, their transitions will eventually become entrained to one another through stochastic synchrony.
q-bio.NC
q-bio
Journal of Mathematical Biology manuscript No. (will be inserted by the editor) Entrainment in up and down states of neural populations: non-smooth and stochastic models Zachary McCleney · Zachary P. Kilpatrick 5 1 0 2 r p A 3 2 ] . C N o i b - q [ 1 v 0 9 2 6 0 . 4 0 5 1 : v i X r a Received: date / Accepted: date Abstract We study the impact of noise on a neural population rate model of up and down states. Up and down states are typically observed in neuronal networks as a slow oscillation, where the population switches between high and low fir- ing rates (Sanchez-Vives and McCormick, 2000). A neural population model with spike rate adaptation is used to model such slow oscillations, and the timescale of adaptation determines the oscillation period. Furthermore, the period depends non-monotonically on the background tonic input driving the population, hav- ing long periods for very weak and very strong stimuli. Using both linearization and fast-slow timescale separation methods, we can compute the phase sensitivity function of the slow oscillation. We find that the phase response is most strongly impacted by perturbations to the adaptation variable. Phase sensitivity functions can then be utilized to quantify the impact of noise on oscillating populations. Noise alters the period of oscillations by speeding up the rate of transition be- tween the up and down states. When common noise is presented to two distinct populations, their transitions will eventually become entrained to one another through stochastic synchrony. Keywords neural adaptation · phase sensitivity function · stochastic synchrony 1 Introduction Cortical networks can generating a wide variety oscillatory rhythms with frequen- cies spanning five orders of magnitude (Buzs´aki and Draguhn, 2004). Slow os- cillatory activity (0.1-1Hz) has been observed in vivo during decreased periods of alertness, such as slow wave sleep and anesthesia (Steriade et al., 1993). Fur- thermore, such activity can be produced in vitro when bathing cortical slices in This work was supported by NSF DMS-1311755. Z. McCleney · Z.P. Kilpatrick Department of Mathematics, University of Houston, Houston TX 77204 E-mail: [email protected] Z. McCleney E-mail: [email protected] 2 McCleney, Kilpatrick a medium with typical extracellular ion concentrations (Sanchez-Vives and Mc- Cormick, 2000). A key feature of these slow oscillations is that they tend to be an alternating sequence of two bistable states, referred to as the up and down states. Up states in networks are characterized by high levels of firing activity, due to depolarization in single cells. Down states in networks typically appear quiescent, due to hyperpolarization in single cells. There is strong evidence that up states are neural circuit attractors, that emerge due to synaptic feedback (Cossart et al., 2003). This suggests up states may be spontaneous remnants of stimulus-induced persistent states utilized for working memory (Wang, 2001) and other network computations (Major and Tank, 2004). Several different cellular and synaptic mechanisms have been suggested to un- derlie the transitions between up and down states. One possibility is that the network is recurrently coupled with excitation, stabilizing both a quiescent and active state (Amit and Brunel, 1997; Renart et al., 2007). Fluctuations due to prob- abilistic synapses, channel noise, and randomness in network connectivity can then lead to spontaneous transitions between the quiescent and active state (Bressloff, 2010; Litwin-Kumar and Doiron, 2012). Alternatively, switches between low and high activity states may arise by some underlying systematic slow process. For in- stance, it has been shown that competition between recurrent excitation and the negative feedback produced by activity-dependent synaptic depression can lead to slow oscillations in firing rate whose timescale is set by the depression timescale (Bart et al., 2005; Kilpatrick and Bressloff, 2010). Excitatory-inhibitory networks with facilitation can produce slow oscillations, due to the slow facilitation of feed- back inhibition that terminates the up state, the down state is then rekindled due to positive feedback from recurrent excitation (Melamed et al., 2008). These neural mechanisms utilize dynamic changes in the strength of neural architecture. However, Compte et al. (2003) proposed that single cell mechanisms can also shift network states between up and down states. The up state is maintained by strong recurrent excitation balanced by inhibition, and transitions to the down state occur due to a slow adaptation current. Once in the down state, the adaptation current is inactivated, and excitation reinitiates the up state. A similar mechanism has been utilized in models of perceptual rivalry, where dominance switches between two mutually inhibiting populations are due to the build up of a rate-based adaptation current (Laing and Chow, 2002; Moreno-Bote et al., 2007). In this paper, we utilize a rate-based model of an excitatory network with spike rate adaptation to explore the impact that noise perturbations have upon the rel- ative phase and duration of slow oscillations. We find that, as in the spiking model studied by Compte et al. (2003), the interplay between recurrent excitation and adaptation produces a slow oscillation in the firing rate of the network. In fact, for slow timescale adaptation currents, the oscillations evolve as fast switches between a low and high activity state, stable fixed points of the adaptation-free system. Since the timescale and slow dynamics of the oscillation are set by the adaptation current, we mainly focus on the impact of perturbation to the adaptation variable in our model. As we will show, perturbations of the activity variable have much lower impact on the oscillation phase. Introducing noise into the adaptation vari- able of the population model leads to a speeding up of the slow oscillation, due to early switching between the low and high activity state. Another remarkable feature of slow oscillations, observed during slow-wave sleep and anesthesia, is that the up and down states tend to be synchronized Population models of up and down states 3 across different regions of cortex and thalamus (Steriade et al., 1993; Massimini et al., 2004). Specifically, both the up and down states start near synchronously in cells located up to 12mm apart (Volgushev et al., 2006). Such remarkable co- herence between distant network activity cannot be accomplished by single cell mechanisms, but require either long range network connectivity or some external signal forcing entrainment (Traub et al., 1996; Smeal et al., 2010). Activity tends to originate from several different foci in the network, quickly spreading across the rest of the network on a timescale orders of magnitude faster than the oscil- lation itself (Compte et al., 2003; Massimini et al., 2004). The fact that the onset of quiescence is fast and well synchronized means there must be either a rapid relay signal between all foci or there is some global signal cueing the down state. Rather than suggest a disynaptic relay, using long range excitation acting on local inhibition, we suggest that background noise can serve as a synchronizing global signal (Ermentrout et al., 2008). Noisy but correlated inputs have been shown to be capable of synchronizing uncoupled populations of phase oscillators (Teramae and Tanaka, 2004) as well as experimentally recorded cells in vitro (Gal´an et al., 2006). Here we will show correlated noise is a viable mechanisms for coordinating slow oscillations in distinct uncoupled neural populations. The paper is organized as follows. We introduce the neural population model in section 2, indicating the way external noise is incorporated into the model. In section 3, we demonstrate the periodic solutions that emerge in the noise-free model, demonstrating it is possible to derive analytical expressions for the oscil- lation period in the case of steep firing rate functions. Then, in section 4 we show how to derive phase sensitivity functions that describe how external perturbations to the periodic solution impact the asymptotic phase of the oscillation. As demon- strated, the impact of perturbations to the adaptation variable is much stronger than activity variable perturbations, especially for longer adaptation timescales. Thus, our studies of the impact of noise mainly focus on the effects of fluctuations in the adaptation variable. We find, in section 5, that adding noise to the adap- tation variable leads to up and down state durations that are shorter and more balanced, so that the up and down state last for similar lengths of time. In sec- tion 6, we demonstrate that slow oscillations in distinct populations can become entrained to one another when both populations are forced by the same common noise signal. This phenomenon is robust to the introduction of independent noise in each population, as we show in section 7. 2 Adaptive neural populations: deterministic and stochastic models We begin by describing the models we will use to explore the impact of external perturbations on slow oscillations. Motivated by Compte et al. (2003), we will focus on a neural population model with spike rate adaptation, akin to mutual inhibitory models used to study perceptual rivalry (Laing and Chow, 2002; Moreno-Bote et al., 2007). Single population model. In a single population, neural activity u(t) receives negative feedback due to a subtractive spike rate adaptation term (Benda and 4 Herz, 2003) McCleney, Kilpatrick u(t) = −u(t) + f (αu(t) − a(t) + I), τ a(t) = −a(t) + φu(t). (1a) (1b) Here, u represents the mean firing rate of the neural population with excitatory connection strength α. The negative feedback variable a is spike frequency adap- tation with strength φ and time constant τ . For some of our analysis we will utilize the assumption τ (cid:29) 1, based on the fact that many form for spike rate adaptation tend to be much slower than neural membrane time constants (Benda and Herz, 2003). The constant tonic drive I initiates the high firing rate (up) state, and slow adaptation eventually attenuates activity to a low firing rate (down) state. Weak but positive drive I > 0 is meant to model the presence of low spiking threshold cells that spontaneously fire, utilized as a mechanism for initiating the up state in Compte et al. (2003). The firing rate function f is monotone and saturating function such as the sigmoid f (x) = 1 1 + e−γx . (2) Commonly, in studies of neural field models, the high gain limit of (2) is taken to yield the Heaviside firing rate function (Amari, 1977; Laing and Chow, 2002) (cid:26) 1 : x ≥ 0, 0 : x < 0, H(x) = (3) which often allows for a more straightforward analytical study of model dynamics. We exploit this fact extensively in our study. Nonetheless, we have also carried out many numerical simulations of the model for a smooth firing rate function (2), and they correspond to the results we presentfor sufficiently high gain. Note, this form of adaptation is often referred to as subtractive negative feedback, as current is subtracted from the population input. Alternative models of slow neural population oscillations have employed short term synaptic depression (Tabak et al., 2000; Bart et al., 2005; Kilpatrick and Bressloff, 2010), a form of divisive negative feedback. A primary concern of this work is the response of (1) to external perturbations, acting on the activity u and adaptation a variables. To do so, we will use both an exact method and a linearization to identify the phase response curve of the limit cycle solutions to (1). Understanding the susceptibility of limit cycles (1) to inputs will help us understand ways in which noise will influence the frequency and regularity of oscillations. Stochastic single population model. Following our analysis of the noise- free system, we will consider how fluctuations influence oscillatory solutions to (1). To do so, we will employ the following Langevin equation for (1) forced by white noise du(t) = [−u(t) + f (αu(t) − a(t) + I)] dt + dξu(t) da(t) = [−a(t) + φu(t)] dt/τ + dξa(t), (4a) (4b) where we have introduced the independent Gaussian white noise processes ξu(t) and ξa(t) with zero mean (cid:104)ξu(t)(cid:105) = (cid:104)ξa(t)(cid:105) = 0 and variances (cid:104)ξu(t)2(cid:105) = σ2 ut Population models of up and down states 5 and (cid:104)ξa(t)2(cid:105) = σ2 at. Extending our results concerning the phase response curve, we will explore how noise forcing impacts the statistics of the resulting stochastic oscillations in (4). In particular, since we find noise tends to impact the phase of the oscillation more strongly when applied to the adaptation variable, we will tend to focus on the case ξu ≡ 0. Stochastic dual population model. Finally, we will focus on how correla- tions in noise-forcing impact the coherence of two distinct uncoupled populations du1 = [−u1(t) + f (αu1(t) − a1(t) + I)] dt + dξu da1 = [−a1(t) + u1(t)] dt/τ + dξa du2 = [−u2(t) + f (αu2(t) − a2(t) + I)] dt + dξu da2 = [−a2(t) + u2(t)] dt/τ + dξa. (5a) (5b) (5c) (5d) Thus, the system (5) describes the dynamics of two distinct neural populations u1 and u2, with inputs I. Our main interest lies in the impact the noise terms have upon the phase relationship between the two systems' states. In this version of the model, noise to the activity variables ξu is totally correlated, as is noise to the adaptation variables ξa. Thus, all means are zero and (cid:104)ξ2 u = D11t. Furthermore, (cid:104)ξ2 at = D22t. For this study, we assume there are no correlations between activity and adaptation noise, so (cid:104)ξu(t)ξa(t)(cid:105) = 0. A more general version of the model (5) would consider the possibility of independent noise in each population u(t)(cid:105) = σ2 a(t)(cid:105) = σ2 1 − χ2 udξu1 1 − χ2 du1 = [−u1(t) + f (αu1(t) − a1(t) + I)] dt + χudξuc + da1 = [−a1(t) + u1(t)] dt/τ + χadξac + adξa1 du2 = [−u2(t) + f (αu2(t) − a2(t) + I)] dt + χudξuc + da2 = [−a2(t) + u2(t)] dt/τ + χadξac + adξa2. Noise terms all have zero mean and variances defined (cid:104)ξ2 ujt = Dujt and (cid:104)ξ2 aj(t)(cid:105) = σ2 ajt = Dajt (j = 1, 2, c). To ease calculations, we take Du1 = Du2 ≡ Dul = σ2 a. The degree of noise correlation between populations is controlled by the parameters χu and χa, so in the limit χu,a → 1, the model (6) becomes (5). u and Da1 = Da2 ≡ Dal = σ2 uj(t)(cid:105) = σ2 1 − χ2 (6a) (6b) 1 − χ2 udξu2 (6c) (6d) (cid:112) (cid:112) (cid:112) (cid:112) 3 Periodic solutions of a single population We begin by studying periodic solutions of the single population system (1), as demonstrated in Fig. 1A. First, we note that for firing rate functions f with finite gain, we can identify the emergence of oscillations by analyzing the stability of the equilibria of (1). That is, we assume ( u, a) = (0, 0), so the system becomes ¯u = f (α¯u − ¯a + I), ¯a = φ¯u, which can be reduced to the single equation ¯u = f ((α − φ)¯u + I) = g(¯u). (7) 6 McCleney, Kilpatrick Fig. 1 Single adapting neural population (1) generates slow oscillations. (A) Numerical simu- lation of (1) for adaptation timescale τ = 100 (1s) and input I = 0.2. (B) Partitioning of (τ, I) parameter space shows the range of inputs I leading to oscillations expands as the adaptation timescale τ is increased, according to (9). (C,D) Bifurcation diagram showing supercritical (I−H ) and subcritical (I+H ) Hopf bifurcations that arise as the input is increased for (C) τ = 10 and (D) τ = 100. Firing rate function is sigmoidal (2). Other parameters are φ = 1, α = 0.5, and γ = 15 Roots of (7), defining fixed points of (1) are plotted as a function of the input I in Fig. 1C,D. Utilizing the sigmoidal firing rate function f given by (2), we can show that there will be a single fixed point as long as φ > α. In this case, we can compute dg(¯u) d¯u = −(φ − α)f (cid:48) ((α − φ)¯u + I) = − (φ − α)e−γ((α−φ)¯u+I) (cid:0)1 + e−γ((α−φ)¯u+I)(cid:1)2 < 0. Since ¯u is monotone increasing, then ¯u − g(¯u) is monotone increasing. Further, noting lim¯u→±∞ [¯u − g(¯u)] = ±∞, it is clear ¯u− g(¯u) crosses zero once, so (7) has a single root when φ > α. Stability of this equilibrium is given by the eigenvalues of the associated Jacobian (cid:18)−1 + αf(cid:48)((α − φ)¯u + I) −f(cid:48)((α − φ)¯u + I) (cid:19) −1/τ . J(¯u, ¯a) = φ/τ Population models of up and down states 7 We note that the sigmoid (2) satisfies the Ricatti equation f(cid:48) = γf (1 − f ), so we can use (7) to write (cid:18)−1 + αγ ¯u(1 − ¯u) −γ ¯u(1 − ¯u) (cid:19) φ/τ −1/τ . J(¯u, ¯a) = Oscillations arise when stable spiral equilibria destabilize through a Hopf bifurca- tion. Hopf bifurcations will occur when complex eigenvalues associated with fixed points (¯u, ¯a) cross from the left to the right half plane. We can require this with the pair of expression: tr(J) = 0 and tr(J)2 < 4det(J). Thus, a necessary condition for the Hopf bifurcation point is that the equilibrium value ¯u satisfy αγ ¯u(1 − ¯u) = 1 + 1/τ. Solving this for ¯u yields ¯u±H = 1 2 (cid:104) 1 ±(cid:112)1 − 4χ (cid:105) , χ = 1 + 1/τ αγ . (8) Thus, Hopf bifurcations will only occur when the timescale of adaptation is suffi- ciently large τ > [αγ/4 − 1] −1. Plugging the formula (8) back into the fixed point equation (7) and solving for the input I, we can parameterize Hopf bifurcation curves based upon the equation (cid:20) (cid:21) I±H = 1 γ ln ¯u±H 1 − ¯u±H − (α − φ)¯u±H , (9) (10) along with the additional condition tr(J)2 < 4det(J) which becomes 4 τ 2 < 4φ ατ 2 + 4φ ατ , which will always hold as long as φ > α. We partition the parameter space (τ, I) using our formula for the Hopf curve (9) in Fig. 1B. As demonstrated, there tend to be either two or zero Hopf bifurcation points for a given timescale τ , and the coalescence of the two Hopf points is given by the point where τ = [αγ/4 − 1] −1. In the limit of slow adaptation τ (cid:29) 1, we can separate the timescales of the activity u and adaptation a variables, finding u will equilibrate according to the equation u(t) = f (αu(t) − a(t) + I), and subsequently a will slowly evolve according to the equation a(t) = [φu(t) − a] /τ. (11) (12) We always have an implicit formula for u(t) in terms of a(t), so the dynamics will tend to slowly evolve along the direction of the a variable. This demonstrates why periodic solutions to (1) are comprised of a slow rise and decay phase of a, punctuated by fast excursions in the activity variable u. In general, it is not straightforward to analytically treat the pair of equations (11) and (12), but we will show how computing solutions of the singular system becomes straightforward when we take the high gain limit γ → ∞. 8 McCleney, Kilpatrick Fig. 2 Analytical approximations to periodic solutions of (1) with a Heaviside firing rate func- tion (3). (A) Numerical simulation (solid lines) of the periodic solution is well approximated by the analytical approximation (dashed lines) given by (15) when I = 0.2 and τ = 100. (B) The period of the oscillation T computed from numerical simulations (dots) is accurately ap- proximated by the analytical formula (solid lines) given by (14). Other parameters are α = 0.5 and φ = 1. Having established the existence of periodic solutions to (1) in the case of sigmoid firing rates (2), we now explore the system in the high gain limit γ → ∞ whereby the firing rate function becomes a Heaviside (3). In this case, fixed points (¯u, ¯a) satisfy the equations ¯u = H((α − φ)¯u + I) = (cid:26) 1 : ¯u < I/(φ − α), 0 : ¯u > I/(φ − α), and ¯a = φ¯u. Thus, assuming φ > α, then ¯u = 0 when I < 0 and ¯u = 1 when I > (φ − α). In both cases, the fixed points are linearly stable. When 0 < I < (φ − α), there are no fixed points and we expect to find oscillatory solutions. Assuming τ (cid:29) 1, we can exploit a separation of timescales to identify the shape and period of these limit cycles. To begin, we note that on fast timescales u(t) = −u(t) + H(I + αu(t) − a0), where a0 is a quasi steady state. On slow timescales on the order of τ , then u(t) quickly equilibrates and u(t) = H(I + αu(t) − a(t)), τ a(t) = −a(t) + φu(t). (13a) (13b) Periodic solutions to (1) must obey the condition (u(t), a(t)) = (u(t + nT ), a(t + nT )) for t ∈ [0, T ] and n ∈ Z, so we focus on the domain t ∈ [0, T ]. Examining (13), we can see oscillations in (1) involve switches between u(t) ≈ 1 and u(t) ≈ 0. We translate time so that u(t) ≈ 1 on t ∈ [0, T1) and u(t) ≈ 0 on t ∈ [T1, T ). Subsequently, this means for t ∈ [0, T1] the system (13) becomes u ≡ 1 and τ a = −a + φ so a(t) = φ − (φ − I)e−t/τ . We know a(0) = I because u(0−) ≡ 0 in (13), and the argument of H(x) must have crossed zero at t = 0. In a similar way, we find on t ∈ [T1, T ) that u ≡ 0 and a(t) = (I + α)e−(t−T1)/τ . Using the Population models of up and down states 9 conditions a(T1) = I +α and a(T ) = I, we find that the rise time of the adaptation variable (or the duration of the up state) is T1 = τ ln (cid:21) , φ − α − I (cid:20) φ − I (cid:21) (cid:20) I + α (cid:20) (I + α)(φ − I) I T2 = τ ln , and the decay time (or the duration of the down state) is and the total period of the oscillation is (cid:21) . (14) Thus, approximate periodic solutions to (1) in the case of a Heaviside firing rate (3) take the form T = τ ln I(φ − α − I) (cid:26) 1 : t ∈ [0, T1), (cid:26) φ − (φ − I)e−t/τ 0 : t ∈ [T1, T ], u(t) = a(t) = : t ∈ [0, T1), (I + α)e−(t−T1)/τ : t ∈ [T1, T ]. (15a) (15b) We demonstrate the accuracy of the approximation (15) in Fig. 2A. Furthermore, we show that relationship between the period T and model parameters is well captured by the formula (14). Notice there is a non-monotonic relationship between the period T and the input I. We can understand this further by noting that the rise time T1 of the adaptation variable a increases monotonically with input dT1 dI = τ α (φ − I)(φ − α − I) > 0, when 0 < I < (φ − α). Furthermore, the decay time T2 of the adaptation variable a decreases monotonically with input dT2 dI = − τ α I(I + α) < 0, when 0 < I < (φ − α). Thus, as I → 0+, the slow oscillation's period T is dominated by very long decay times T2 (cid:29) 1 and as I → (φ− α)−, it is dominated by very long rise times T1 (cid:29) 1. We can identify the minimal period as a function of the input I by finding the critical point of T (I). To do so, we differentiate and simplify = − dT dI τ αφ(2I − (φ − α)) I(I + α)(φ − I)(φ − α − I) , so the critical point of T (I) is Icrit = (φ− α)/2, which corresponds to the minimal value of the period Tmin(I) = 2τ ln [(φ + α)/(φ − α)] as pictured in Fig. 2B. 10 McCleney, Kilpatrick Fig. 3 (A, B, C) Periodic solution (u, a) and (D, E, F) phase sensitivity function (Zu, Za) of (1) plotted as a function of phase θ = t/T for a sigmoidal firing rate function (2). (A,D) For shorter adaptation timescale τ = 10 and input I = 0.2, the activity variable u has a more rounded trajectory, so perturbations to activity influence the oscillation phase more (note size of lobes in on Zu in (D). (B,E) As the adaptation timescale is increased to τ = 100, with I = 0.2, the influence of perturbations to the activity variable decrease (compare lobes of Zu to those in (D). Perturbations of the adaptation variable influence the phase more strongly as shown by the change in the relative amplitude of Za. (C,F) Increasing the input I = 0.4, with τ = 10, increases the relative duration of the rise time of a. As a result, there is a wider region where perturbations to a advance the phase. Other parameters are α = 0.5, φ = 1, and γ = 15. 4 Phase response curves We can further understand the dynamics of the slow oscillations in (1) by comput- ing phase response curves for both the case of a sigmoidal firing rate (2) and the Heaviside firing rate (3). As we will show, perturbations of the activity variable u have decreasing impact as the timescale of adaptation τ and the gain γ of the firing rate are increased. Perturbations of the adaptation variable a tend to dominate the resulting dynamics, as it is the evolution of this slow variable that primarily determines the phase of the oscillation. (cid:19) L To begin, we derive a general formula that linearly approximates the influence of small perturbations on limit cycle solutions (u0(t), a0(t)) to (1). Essentially, we utilize the fact that solutions Z(t) to the adjoint equation associated with lineariza- tion about the limit cycle solution (u0(t), a0(t)) provide a complete description of how infinitesimal perturbations of the limit cycle impact its phase (Ermentrout, 1996; Brown et al., 2004). To start, we note that (cid:18) u1 + u1 − αf(cid:48)(αu0 − a0 + I)u1 + f(cid:48)(αu0 − a0 + I)a1, (cid:18) u1 product on T -periodic functions in R2 as (cid:104)F (t), G(t)(cid:105) =(cid:82) T is the linearization of (1) about the limit cycle (u0(t), a0(t)). Defining the inner 0 F (t) · G(t)dt, we can find the adjoint operator L∗ by noting it satisfies (cid:104)F,LG(cid:105) = (cid:104)L∗F, G(cid:105) for all L2 a1 − φu1/τ + a1/τ (cid:18) 0 (cid:19) , 0 = a1 (cid:19) = Population models of up and down states integrable vector functions F, G. We can then compute L∗(cid:18) v b (cid:19) = (cid:18)− v + v − αf(cid:48)(αu0 − a0 + I)v − φb/τ −b + f(cid:48)(αu0 − a0 + I)v + b/τ (cid:19) . 11 (16) It can be shown that the null space of L∗ describes the response of the phase of the limit cycle (u0(t), a0(t)) to infinitesimal perturbations (Brown et al., 2004). Note that if (u0(t), a0(t)) is a stable limit cycle then the nullspace of L is spanned 0(t), a(cid:48) by scalar multiples of (u(cid:48) 0(t)). Furthermore, appropriate normalization re- quires that Z(t) · (u(cid:48) 0(t)) = 1 along with L∗Z = 0 (Ermentrout, 1996). To 0(t), a(cid:48) numerically compute Z(t) = (Zu(t), Za(t)), we thus integrate the system Zu(t) = Zu(t) − αf Za(t) = f (cid:48) (cid:48) (αu0(t) − a0(t) + I)Zu(t) − φZa(t)/τ, (17a) (αu0(t) − a0(t) + I)Zu(t) − Za(t)/τ, (17b) backward in time, taking the long time limit to find (Zu(t), Za(t)) on t ∈ [0, T ], and normalizing (cid:104)(Zu(t), Za(t)), (u(cid:48) 0(t))(cid:105) = 1 by rescaling appropriately. We demonstrate this result in Fig. 3, showing the relationship between the shape and relative amplitude of the phase sensitivity functions (Zu, Za) and the parameters. Notably, perturbing the activity variable u become less influential as the timescale of adaptation τ is increased (Zu). Furthermore, there is a sharper transition be- tween phase advance and phase delay region of the adaptation phase response (Za) for larger timescales τ . 0(t), a(cid:48) In addition to a general formula for the phase sensitivity functions (Zu(t), Za(t)), we can derive an amplitude dependent formula for the response of limit cycle so- lutions (u0(t), a0(t)) of (1) with a Heaviside firing rate (3), assuming τ (cid:29) 1. In this case, we utilize the formula for the period (14) and limit cycle (15), derived using a separation of timescales assumption. Then, we can compute the change to the variables (u, a) as a result of a perturbation (δu, δa), which we denote (cid:55)−→ (u0(t), a0(t)). We are primarily interested in how the relative (u0(t), a0(t)) time in the limit cycle is altered by a perturbation δu - how much closer or further the limit cycle is to the end of the period T after being perturbed. We can readily determine this by first inverting the formula we have for (u0(t), a0(t)), given by (15), to see how this value determines the time t0 along the limit cycle (δu,δa) t0(u0, a0) = τ ln [(φ − I)/(φ − a0)] : u0 = 1, τ ln [(φ − I)(I + α)/a0/(φ − α − I)] : u0 = 0. (18) (cid:26) Using this formula, we can now map the value (u0, a0) to an associated updated relative time t0 along the oscillation. Here, we decompose the impact of perturbations to the u and a variables. We begin by studying the impact of perturbations δu to the activity variable u. We can directly compute u0(t) = H(I + α [u0(t) + δu] − a0(t)). Thus, the singular system (13) will be unaffected by such perturbations if sgn(I + α[u + δu]− a) = sgn(I + αu− a). This is related to the flatness of the susceptibility function Zu over much of the time domain in Fig. 3D,E,F. However, if sgn(I + 12 McCleney, Kilpatrick Fig. 4 Phase response curves of the fast-slow timescale separated system τ (cid:29) 1. (A,B) Amplitude δu- and δa-dependent phase response curves Gu(θ, δu) and Ga(θ, δu) characterizing phase advances/delays resulting from perturbation of neural activity u and adaptation a. We compare analytical formulae (solid lines) to numerically computed PRCs (dashed lines). (C) Phase response curve associated with perturbations of the adaptation variable a in the small amplitude 0 < δa (cid:28) 1 limit. We compare the large amplitude formula (solid line) determined by (24) to the linear approximation (dotted line) given by (25) to numerical computations (dashed line). α[u + δu]− a) (cid:54)= sgn(I + αu− a), then u0(t) = 1− u0(t), as detailed in the following piecewise smooth map: u0(t) = 0 (cid:55)→ u0(t) = 1 : δu > −(I − a0(t))/α > 0, u0(t) = 0 (cid:55)→ u0(t) = 0 : δu < −(I − a0(t))/α > 0, u0(t) = 1 (cid:55)→ u0(t) = 0 : −δu < −(I + α − a0(t))/α < 0, u0(t) = 1 (cid:55)→ u0(t) = 1 : −δu > −(I + α − a0(t))/α < 0, where (u0(t), a0(t)) are defined by (15). The formula (18) can then be utilized to compute the updated relative time t0 := t0(u0, a0), finding  τ ln [(φ − I)(I + α)/a0/(φ − α − I)] T + τ ln [(φ − I)/(φ − a0)] t0(u0, a0) t0 = : δu > (I + α − a0)/α > 0, u0 = 1 : −δu > (a0 − I)/α > 0, u0 = 0 : otherwise, (19) where a0 = φ − (φ − I)e−t0/τ if u0 = 1 and a0 = (I + α)e−(t0−T1)/τ if u0 = 0. We can refer to the function t0/T , where t0 is defined by (19), as the phase transition curve for u perturbations. Thus, the function Gu(θ, δu) = (t0 − t0)/T will be the phase response curve, where θ = t0/T , and phase advances occur for positive values and phase delays occur for negative values. We plot the function Gu(θ, δu) in Fig. 4A for different values of δu, demonstrating the nontrivial dependence on the perturbation amplitude is not simply a rescaling but an expansion of the non-zero phase shift region. Due to the singular nature of the fast-slow limit cycle (15), the size of the phase perturbation has a piecewise constant dependence on the amplitude of the u perturbation. Note, this formulation allows us to quantify phase shifts that would not be captured by a perturbative theory for phase sensitivity functions, as computed for the general system in (17). For perturbations δa of the adaptation variable a, there is a more graded de- pendence of the phase advance/delay amplitude on the perturbation amplitude δa. We expect this, as it was a property we observed in Za as we varied parameters in Population models of up and down states 13 Fig. 3. We can partition the limit cycle (u0(t), a0(t)) into four different regions: two advance/delay regions of exponential saturation and two early threshold crossings. First, note if u0(t) = 1 and a0(t) + δa < I + α, then a0(t) = φ − (φ − I)e (20) so t0 = T1−tw with tw = τ ln [(φ − a0 − δa)/(φ − I − α)], but if a0(t)+δa > I +α, then −t/τ + δa, u0(t) = 1, u0(t) = 0, a0(t) = φ − (φ − I)e −t/τ + δa. (21) Determining the relative time of the perturbed variables (u0(t), a0(t)) in (20) is straightforward using the mapping (18). However, to determine the relative time described by (21), we compute the time, after the perturbation, until a0(t) = I +α, which will be tw = τ ln [(a0 + δa)/(I + α)], so t0 = T1− tw. Second, note if u0(t) = 0 and a0(t) + δa > I, then u0(t) = 0, (22) so t0 = T − tw with tw = τ ln [(a0 + δ0)/I], but if a0(t) + δa < I, so that it is necessary that δa < 0, then a0(t) = (I + α)e −(t−T1)/τ + δa, u0(t) = 1, (23) In the case of (23), we note that tw = τ ln [(φ − a0 − δa)/(φ − I)], so t0 = T − tw. Combining our results, we find we can map the relative time to the perturbed relative time as a0(t) = (I + α)e −(t−T1)/τ + δa. T1 − τ ln [(φ − a0 − δa)/(φ − I − α)] T1 − τ ln [(a0 + δa)/(I + α)] T − τ ln [(a0 + δa)/I] T − τ ln [(φ − a0 − δa)/(φ − I)] : δa < I + α − a0, u0 = 1, : δa > I + α − a0, u0 = 1, : δa > I − a0, u0 = 0, : δa < I − a0, u0 = 0, (24)  t0 = where a0 = φ − (φ − I)e−t0/τ if u0 = 1 and a0 = (I + α)e−(t0−T1)/τ if u0 = 0. Again, we have a phase transition curve given by the function t0/T and phase response curve given by Ga(θ, δa) = (t0 − t0)/T , where θ = t0/T . As opposed to the case of u perturbations, the phase perturbation here depends smoothly on the amplitude of the a perturbation δa. Furthermore, we can obtain a perturbative description of the phase response curve for the singular system (13) in two ways: (a) Taylor expand the amplitude- dependent phase response curve expressions defined by (19) and (24) and truncate to linear order or (b) solving the adjoint equation (17) in the case of a Heaviside firing rate (3) and long adaptation timescale τ (cid:29) 1. We begin with the first derivation, which simply requires differentiating (19) to demonstrate that the in- finitesimal phase response curve (iPRC) associated with perturbations of the u variable is zero almost everywhere. However, differentiating (24) reveals that the iPRC associated with perturbations of the adaptation variable a is given by the piecewise smooth function Za(t) = τ T (φ − I)s − τ T (I + α) : t ∈ (0, T1), et/τ e(t−T1)/τ : t ∈ (T1, T ). (25)  14 McCleney, Kilpatrick Furthermore, note we could derive the same result by solving the adjoint equations (17) in the case of Heaviside firing rate (3), so that Zu(t) = −Zu(t) + αδ(αu0(t) − a0(t) + I)Zu(t) + φZa(t)/τ, Za(t) = −δ(αu0(t) − a0(t) + I)Zu(t) + Za(t)/τ. (26b) Note, we have reversed time t (cid:55)→ −t, so we can simply solve the system forward. Furthermore, we can use the identity (26a) δ(αu0(t) − a0(t) + I) = δ(t) u(cid:48)(0) − a(cid:48)(0) + δ(t − T1) u(cid:48)(T1) − a(cid:48)(T1) . (27) Utilizing the separation of timescales, τ (cid:29) 1, we find that almost everywhere (except where t = 0, T1, T ), we have that (26) becomes the system Zu(t) = −Zu(t), τ Za(t) = Za(t). (28) As before Zu(t) will be zero almost everywhere, whereas Za(t) = A(t)et/τ , where A(t) is a piecewise constant function taking two different values on t ∈ (0, T1) and t ∈ (T1, T ), determined by considering the δ distribution terms. This indicates how one would derive the formula (25) using the adjoint equations (26). Note, in previous work (Jayasuriya and Kilpatrick, 2012), we explored the en- trainment of slowly adapting populations to external forcing, comprised of smooth and non-smooth inputs to the system (1). In the next section, we explore the im- pact of external noise forcing on the slow oscillations of (1), subsequently demon- strating that noise can be utilized to entrain the up and down states of two distinct networks. 5 Impact of noise on the timing of up/down states We now study the effects of noise on the duration of up and down states of the single population model (1). Switches between high and low firing rate states can occur at irregular intervals (Sanchez-Vives and McCormick, 2000), suggesting internal or external sources of noise determine state changes. This section focuses on how noise can reshape the mean duration of up and down residence times. Due to our findings in the previous sections, we focus on noise applied to the adaptation variable in this section. As we have shown, very weak perturbations to the neural activity variable have a negligible effect on the phase of oscillations. Analytic calculations are presented for the piecewise smooth system with Heaviside firing rate (3), as accurate approximations of the mean up and down state durations can be computed. Our approach is to derive expressions for the mean first passage times of both the up and down state ( ¯T1 and ¯T2) of the stochastic population model (4). Focusing on adaptation noise allows us to utilize the separation of fast-slow timescales, and recast the pair of equations as a stochastic-hybrid system u(t) = H((αu(t) + I − a(t)), da(t) = [−a(t) + φu(t)] dt/τ + dξa(t), Population models of up and down states 15 Fig. 5 Noise alters the duration of up and down states. (A) Numerical simulation of the stochastically driven population model (4) demonstrates up and down state durations (e.g., T1 and T2) are variable when driven by adaption noise ξa with (cid:104)ξ2 at, σa = 0.01. Switches are determined by the threshold crossings of the adaptation variable a(t) = I and a(t) = I + α. (B) Up/down states become more variable when the noise amplitude σa = 0.02. (C) Mean durations of the up and down state, (cid:104)T1(cid:105) and (cid:104)T2(cid:105), decrease as a function noise amplitude σa. (D) Impact of noise σa on the balance of up to down state durations ¯T1/ ¯T2 as input I is varied. Firing rate is given by the Heaviside function (3). Other parameters are α = 0.5, φ = 1, and τ = 50. a(cid:105) = σ2 where ξa is white noise with mean (cid:104)ξa(cid:105) = 0 and variance (cid:104)ξ2 at. To begin, assume the system has just switched to the up state, so the initial conditions are u(0) = 1 and a(0) = I. Determining the amount of time until a switch to the down state requires we calculate the time T1 until the threshold crossing a(T1) = I + α where a(t) is determined by the stochastic differential equation (SDE) a(cid:105) = σ2 da(t) = [−a(t) + φ] dt/τ + dξa, which is the well-known threshold crossing problem for an Ornstein-Uhlenbeck process Gardiner (2004). The mean ¯T1 of the passage time distribution is thus given by defining the potential V (a) = a2 τ and computing the integral (cid:90) I+α (cid:90) x 2τ − φa e[V (x)−V (y)]/σ2 a dydx. ¯T1 = 1 σ2 a I −∞ Next, note that the duration of the down state T2 will be the amount of time until the threshold crossing a(T2) = I given u(0) = 0 and a(0) = I + α, where a(t) obeys the SDE da(t) = [−a(t)] dt/τ + dξa(t). 16 McCleney, Kilpatrick Again, defining the potential V (a) = a2 2τ , we can compute e[V (x)−V (y)]/σ2 a dydx. (cid:90) −I (cid:90) x −I−α −∞ ¯T2 = 1 σ2 a We compare the theory we have developed utilizing passage time problems to residence times computed numerically in Fig. 5C. Notice that increasing the noise amplitude tends to shorten both up and down state durations on average, due to early threshold crossings of the variable a(t). Furthermore, we can examine how noise reshapes the relative balance of up versus down state durations. Specifically, we will explore how the relative fraction of time the up state persists ¯T1/( ¯T1 + ¯T2) changes with noise intensity σa and input I. First, notice that, in the absence of noise the ratio ln [(φ − I)/(φ − α − I)] T1 T1 + T2 = ln [(I + α)(φ − I)/I(φ − α − I)] . (29) The up and down state have equal duration when T1/(T1 + T2) = 1/2, or when the input I = (φ − α)/2, as shown in Fig. 5D. Interestingly, this is the precise input value at which the period obtains a minimum, as we demonstrated in section 3. Along with our plot of (29) in the noise-free case (σa = 0), we also study the impact of noise on this measure of up-down state balance. Noise leads to up and down state durations becoming more similar, so the ratio (29) of the means ¯T1 and ¯T2 flattens as a function of the input I. This is due to the fact that long durations, wherein the variable a(t) occupies the tail of exponentially saturating functions A0 + A1e−t/τ , are shortened by early threshold crossings due to the external noise forcing. 6 Synchronizing two uncoupled populations Now we demonstrate that common noise can synchronize the up and down states of two distinct and uncoupled populations. We begin with the case of identical noise and then, in section 7, relax these assumptions to show that some level of coherence is still possible when each population has an intrinsic and independent source of noise. This is motivated by the observation that the neural Langevin equation derived in the large system-size limit of a neural master equation tends to possess intrinsic noise in each population, in addition to an extrinsic common noise term (Bressloff and Lai, 2011). As we will show, intrinsic noise tends to disrupt the phase synchronization due to extrinsic noise. To begin, we recast the stochastic system (5), describing a pair of adapting noise-driven neural populations, as a pair of phase equations: dθ1(t) = ωdt + Z(θ1(t)) · dξ(t), dθ2(t) = ωdt + Z(θ2(t)) · dξ(t), (30a) (30b) where θ1 and θ2 are the phase of the first and second neural populations. As we demonstrate in Fig. 6A, this introduction of common noise tends to drive the oscillation phases θ1(t) and θ2(t) toward one another. Note that since the governing equations of both populations are the same, then the phase sensitivity Population models of up and down states 17 Fig. 6 Synchronizing slow oscillations in two uncoupled populations described by (5) with sigmoidal firing rate (2). (A) Single realization of the system (5) driven by common noise ξa to the adaptation variable ((cid:104)ξ2 a(cid:105) = ε2t, ε = 0.01) with input I = 0.2 and adaptation timescale τ = 50. Notice that the phase difference ψ(t) = ∆1(t) − ∆2(t) roughly decreases over time. (B) Plot of the log of the phase difference y(t) = ln ψ(t) for several realizations (thin lines) compared with the theory (thick line) of the mean y(0) + λt computed using the Lyapunov exponent (35). (C) Lyapunov exponent λ decreases as a function of the adaptation timescale τ , for I = 0.2. We compare numerical simulations (dots) to theory (solid). (D) Lyapunov exponent λ varies non-monotonically with the strength of the input I. Other parameters are α = 0.5, γ = 15, and φ = 1. function Z(θ) will be the same for both. Furthermore, the synchronized solution θ1(t) = θ2(t) is absorbing -- once the phases synchronize, they remain so. We can analytically calculate the Lyapunov exponent λ of the synchronized state to determine its stability. In particular, we will be interested in how this stability depends on the parameters that shape the dynamics of adaptation. We next convert the pair of Stratonovich differential equations into a equivalent pair of Ito differential equations: (cid:105) (cid:105) dt + Z(θ1(t)) · dξ(t), dt + Z(θ2(t)) · dξ(t), (31a) (31b) (cid:104) (cid:104) dθ1(t) = dθ2(t) = (cid:48) (cid:48) ω + Z ω + Z (θ1(t))T DZ(θ1(t)) (θ2(t))T DZ(θ2(t)) introducing a drift term due to our change in definition of the noise term. Now, to determine stability of the synchronized state θ1(t) = θ2(t), we assume we are infinitesimally close to it. We define ψ(t) = θ1(t) − θ2(t) and require ψ(t) (cid:28) 1. Linearizing the system of Ito differential equations (31) with respect to ψ(t) then 18 yields (cid:20)(cid:16) (cid:48) Z dψ(t) = ψ(t) (cid:17)(cid:48) dt + Z (cid:48) (θ(t)) · dξ (cid:21) , (θ(t))T DZ(θ(t)) (32) McCleney, Kilpatrick where θ(t) obeys either one of the equations in (31). Applying the change of variables y(t) = ln ψ(t), we can rewrite (32) as (cid:48)T DZ dy(t) =(cid:0)Z dt −(cid:16) DZ(cid:1)(cid:48) (cid:48) · dξ(t). (cid:48)(cid:17) dt + Z (33) Z (cid:48) (cid:90) t 0 (cid:90) 1 Notice, on average, the log of the phase difference y(t) tends to decrease over time (Fig. 6B), indicating the phases θ1 and θ2 move toward one another. Subsequently, we can integrate equation (33) to determine the mean drift of y(t) (cid:104)(cid:0)Z (cid:48) (θ(s))DZ(θ(s))(cid:1)(cid:48) −(cid:16) (cid:17)(cid:105) (cid:48)T (θ(s))DZ (cid:48) (θ(s)) Z ds, λ := lim t→∞ so the phase difference ψ(t) will tend to decay grow if the Lyapunov exponent λ < 0 (λ > 0), and the synchronous state will be stable (unstable). Now, utilizing ergodicity of (33), we can compute λ using the ensemble average across realizations of Z(cid:48)(θ(t)) · ξ(t), so (cid:104)(cid:0)Z (cid:48) (θ)DZ(θ)(cid:1)(cid:48) −(cid:16) (cid:48)T (θ)DZ (cid:48) (cid:17)(cid:105) 0 λ = Ps(θ) (34) where Ps(θ) is the steady state distribution of θ. Since noise is weak (Djk (cid:28) 1, j, k = 1, 2), we can approximate the distribution as constant Ps(θ) = 1. Upon applying this to the integrand of (34) and noting the periodicity of Z(θ), we find we can approximate the Lyapunov exponent dθ, (θ) Z λ = − (cid:48)T (θ)DZ (cid:48) Z (θ)dθ. (35) Assuming noise to the activity variable u and adaptation variable a is not cor- related, D will be diagonal. In this case, we can further decompose the phase sensitivity function into its Fourier expansion (cid:90) 1 0 ∞(cid:88) k=0 ∞(cid:88) 2π2k2(cid:104)(cid:16) k=0 Z(θ) = ak sin(2πkθ) + bk cos(2πkθ), where ak = (ak1, ak2)T and bk = (bk1, bk2)T are vectors in R2 so that (cid:48) Z (θ) = 2πk [ak cos(2πkθ) − bk sin(2πkθ)] , and we can expand the terms in (35) to yield ∞(cid:88) λ = − (cid:17) (cid:16) a2 k2 + b2 k2 (cid:17) (cid:105) . D22 a2 k1 + b2 k1 D11 + k=0 Thus, as long as Z(θ) is continuous and non-constant, the Lyapunov exponent λ will be negative, so the synchronous state θ1 = θ2 will be stable. Note, continuity is Population models of up and down states 19 Fig. 7 Stationary density M0(ψ) of the phase difference ψ = θ1 − θ2 for two slowly oscillating neural population driven by both common and independent noise (6). As the degree of noise correlation is decreased from (A) χa = 0.95 to (B) χa = 0.90, the density spreads, but there is still a peak at ψ = 0, the phase-locked state. We focus on noise in the adaptation variable, so σu = 0 and σa = 0.01. Other parameters are α = 0.5, γ = 15, φ = 1, and τ = 20. not satisfied in the case of our singular approximation to Z(θ). We demonstrate the accuracy of our theory (35) in Fig. 6C,D, showing that λ decreases as a function of τ and is non-monotonic in I. Thus, slow oscillations with longer periods are synchronized more quickly, relative to the number of oscillation cycles. Since the Lyapunov exponent has highest amplitude λ for both low and high values of the tonic input I, we also suspect this is related to the period of the oscillation T . 7 Impact of intrinsic noise on stochastic synchronization We now extend our results from the previous section by studying the impact of independent noise in each population. Independent noise is incorporated into the modified model (6). Noting, again there is a periodic solution to the noise-free version of this system, phase-reduction methods can be used to obtain approximate Langevin equations for the phase variables (Nakao et al., 2007) dθ1 = ωdt + Z(θ1(t)) · [dξc(t) + dξ1(t)] , dθ2 = ωdt + Z(θ2(t)) · [dξc(t) + dξ2(t)] , where the noise vectors ξc = (χuξuc, χaξac)T and ξj = ((cid:112)1 − χ2 aξaj)T (j = 1, 2). We can reformulate the pair of Stratonovich differential equations as Ito stochastic differential equations given by the system uξuj,(cid:112)1 − χ2 (36b) (36a) dθ1 = A1(θ)dt + dζ1(θ, t), dθ2 = A2(θ)dt + dζ2(θ, t), where the statistics of the noise terms dζj(θ, t) = Z(θj(t)) ·(cid:2)dξc(t) + dξj(t)(cid:3) (j = (cid:17) 1, 2) are specified by (cid:104)dζj(θ, t)(cid:105) = 0 and (cid:104)dζj(θ, t)dζk(θ, t)(cid:105) = Cjk(θ)dt where (cid:17)(cid:16) (37b) (37a) (cid:16) ac Za(θj) χuD1/2 uc Zu(θk) + χaD1/2 ac Za(θk) Cjk(θ) = χuD1/2 (cid:16)(cid:112) + uc Zu(θj) + χaD1/2 1 − χ2 ul Zu(θj) + uD1/2 (cid:112) (cid:17)2 1 − χ2 aD1/2 al Za(θj) δj,k, (38) 20 McCleney, Kilpatrick separating the impact of correlated and local sources of noise. The drift terms can thus be calculated Aj(θ) = ω + 1 Cjj(θ). The Fokker-Planck equation 4 describing the evolution of the probability density function P (θ, t) of the phases is given ∂ ∂θj = − 2(cid:88) j=1 ∂P ∂t 2(cid:88) 2(cid:88) j=1 k=1 ∂ ∂θj [Aj(θ)P ] + 1 2 ∂2 ∂θj∂θk [Cjk(θ)P ] . (39) Now, we apply a change of variables to the Fokker-Planck equation (39) defined θj = ωt + ϑj. Assuming noise is weak, the function Q(ϑ, t) varies slowly compared with the period of the phase oscillators θj. Thus, we average the drifts Aj(θ) and correlation function Cjk(θ) over a single period. The resulting Fokker-Planck equation is then 2(cid:88) 2(cid:88) j=1 k=1 ∂Q(ϑ, t) ∂t = 1 2 ∂2 ∂ϑj∂ϑk [Bjk(ϑ)Q] where the averaged correlation function is given by the formula Bjk(ϑ) = g(θ1 − θ2) + h(0)δj,k, (cid:90) 1 (cid:104) (cid:105) (cid:48) dθ where the correlation functions are defined (cid:20) ∂2Q ∂ϑ2 1 (cid:21) g(θ) = χ2 uDucZu(θ (cid:48) (cid:48) )Zu(θ + θ) + χ2 aDacZa(θ (cid:48) (cid:48) )Za(θ + θ) and h(θ) = (cid:90) 1 (cid:104) 0 0 (1 − χ2 u)DulZu(θ (cid:48) (cid:48) )Zu(θ + θ) + (1 − χ2 a)DalZa(θ (cid:48) (cid:48) )Za(θ + θ) (cid:105) (cid:48) . dθ We study the relationship between the phases ϑ1 and ϑ2 by substituting the for- mula for the averaged correlation matrix ∂Q(ϑ, t) ∂t = 1 2 [g(0) + h(0)] + ∂2Q ∂ϑ2 2 + ∂2 ∂ϑ1∂ϑ2 [g(ϑ1 − ϑ2)Q] . (40) We can write (40) as a separable equation by employing a change of variables that tracks the average ρ = (ϑ1 + ϑ2)/2 and phase difference ψ = ϑ1− ϑ2 of the original position variables, so ∂U (ρ, t) ∂t ∂M (ψ, t) ∂t = = 1 4 ∂2 [g(0) + g(ψ) + h(0)] ∂2U (ρ, t) ∂ρ2 , ∂ψ2 [g(0) − g(ψ) + h(0)] M (ψ, t). (41a) (41b) Thus, we can solve for the stationary solution of the system (41a) by serving Ut = Mt ≡ 0 and requiring periodic boundary conditions. We find that the stationary distribution of the position average is U0(ρ) = 1. In addition, we can integrate the stationary equation for M (ψ, t) to find M0(ψ) = u [(2 − χ2 σ2 u)gu(0) − χ2 ugu(ψ)] + σ2 a [(2 − χ2 a)ga(0) − χ2 aga(ψ)] m0 , (42) Population models of up and down states 21 where m0 = 1/(cid:82) 1 simplified the expression using Du1 = Du2 ≡ Dul = σ2 σ2 a and defined 0 [g(0) − g(x) + h(0)] (cid:90) 1 −1 dx is a normalization factor and we have u and Da1 = Da2 ≡ Dal = gj(ψ) = Zj(θ)Zj(θ + φ)dθ. 0 When noise to each layer is independent (χu, χa → 0), then M0(ψ) = 1 is constant in space. When noise is totally correlated (χu, χa → 1), then M0(ψ) = δ(φ). The stationary distribution M0(ψ) will broaden as the correlations χu and χa are decreased from unity, with a peak remaining at φ = 0. We demonstrate the accuracy of the formula (42) for the stationary density of the phase difference ψ in Fig. 7, showing that it widens as the level noise correlation is decreased. Again, we focus on the impact of adaptation noise. Thus, even when independent noise is introduced, there is some semblance of synchronization in the phases of two noise-driven neural populations (6). 8 Discussion We have studied the impact of deterministic and stochastic perturbations to a neural population model of slow oscillations. The model was comprised of a sin- gle recurrently coupled excitatory population with negative feedback from a slow adaptive current (Laing and Chow, 2002; Jayasuriya and Kilpatrick, 2012). By examining the phase sensitivity function (Zu, Za), we found that perturbations of the adaptation variable lead to much larger changes in oscillation phase than per- turbations of neural activity. Furthermore, this effect becomes more pronounced as the timescale τ of adaptation is increased. Introducing noise in the model decreases the oscillation period and helps to balance the mean duration of the oscillation's up and down states. When two uncoupled populations receive common noise, their oscillation phases θ1 and θ2 eventually become synchronized, which can be shown by deriving a formula for the Lyapunov exponent of the absorbing state θ1 ≡ θ2 (Teramae and Tanaka, 2004). When independent noise is introduced to each popu- lation, in addition to common noise, the long-term state of the system is described by a probability density for ψ = θ1 − θ2, which peaks at ψ ≡ 0. Our study was motivated by the observation that recurrent cortical networks can spontaneously generate stochastic oscillations between up and down states. Guided by previous work in spiking models (Compte et al., 2003), we explored a rate model of a recurrent excitatory network with slow spike frequency adaptation. One of the open questions about up and down state transitions concerns the degree to which they are generated by noise or by more deterministic mechanisms, such as slow currents or short term plasticity (Cossart et al., 2003). Here, we have provided some characteristic features that emerge as the level of noise responsible for transitions is increased. Similar questions have been probed in the context of models of perceptual rivalry (Moreno-Bote et al., 2007). In addition, we have provided a plausible mechanism whereby the onset of up and down states could be synchronized in distinct networks (Volgushev et al., 2006). There are several potential extensions to this work. For instance, we could ex- amine the impact of long-range connections between networks to see how these 22 McCleney, Kilpatrick interact with common and independent noise to shape the phase coherence of oscillations. Similar studies have been performed in spiking models by Ly and Er- mentrout (2009). Interestingly, shared noise can actually stabilize the anti-phase locked state in this case, even though it is unstable in the absence of noise. Fur- thermore, it is known that coupling spanning long distances can be subject to axonal delays. In spite of this, networks of distantly coupled clusters of cells can still sustain zero-lag synchronized states (Vicente et al., 2008). Thus, we could also explore the impact of delayed coupling, determining how features of phase sensi- tivity function interact with delay to promote in-phase or anti-phase synchronized states. References Amari S (1977). Dynamics of pattern formation in lateral-inhibition type neural fields. Biol Cybern 27:77 -- 87. Amit DJ, Brunel N (1997). Model of global spontaneous activity and local struc- tured activity during delay periods in the cerebral cortex. Cerebral cortex 7:237 -- 252. Bart E, Bao S, Holcman D (2005). Modeling the spontaneous activity of the auditory cortex. Journal of computational neuroscience 19:357 -- 378. Benda J, Herz AVM (2003). A universal model for spike-frequency adaptation. Neural Comput 15:2523 -- 2564. Bressloff PC (2010). Metastable states and quasicycles in a stochastic wilson-cowan model of neuronal population dynamics. Physical Review E 82:051903. Bressloff PC, Lai YM (2011). Stochastic synchronization of neuronal populations with intrinsic and extrinsic noise. The Journal of Mathematical Neuroscience (JMN) 1:1 -- 28. Brown E, Moehlis J, Holmes P (2004). On the phase reduction and response dynamics of neural oscillator populations. Neural computation 16:673 -- 715. Buzs´aki G, Draguhn A (2004). Neuronal oscillations in cortical networks. science 304:1926 -- 1929. Compte A, Sanchez-Vives MV, McCormick DA, Wang XJ (2003). Cellular and network mechanisms of slow oscillatory activity (¡ 1 hz) and wave propagations in a cortical network model. Journal of neurophysiology 89:2707 -- 2725. Cossart R, Aronov D, Yuste R (2003). Attractor dynamics of network up states in the neocortex. Nature 423:283 -- 8. Ermentrout B (1996). Type i membranes, phase resetting curves, and synchrony. Neural computation 8:979 -- 1001. Ermentrout GB, Gal´an RF, Urban NN (2008). Reliability, synchrony and noise. Trends in neurosciences 31:428 -- 434. Gal´an RF, Fourcaud-Trocm´e N, Ermentrout GB, Urban NN (2006). Correlation- induced synchronization of oscillations in olfactory bulb neurons. The Journal of neuroscience 26:3646 -- 3655. Gardiner CW (2004). Handbook of stochastic methods for physics, chemistry, and the natural sciences. Berlin: Springer-Verlag. Jayasuriya S, Kilpatrick ZP (2012). Effects of time-dependent stimuli in a com- petitive neural network model of perceptual rivalry. Bulletin of mathematical biology 74:1396 -- 1426. Population models of up and down states 23 Kilpatrick ZP, Bressloff PC (2010). Effects of synaptic depression and adaptation on spatiotemporal dynamics of an excitatory neuronal network. Physica D: Nonlinear Phenomena 239:547 -- 560. Laing CR, Chow CC (2002). A spiking neuron model for binocular rivalry. J Comput Neurosci 12:39 -- 53. Litwin-Kumar A, Doiron B (2012). Slow dynamics and high variability in balanced cortical networks with clustered connections. Nature neuroscience 15:1498 -- 1505. Ly C, Ermentrout GB (2009). Synchronization dynamics of two coupled neural os- cillators receiving shared and unshared noisy stimuli. Journal of computational neuroscience 26:425 -- 443. Major G, Tank D (2004). Persistent neural activity: prevalence and mechanisms. Current opinion in neurobiology 14:675 -- 684. Massimini M, Huber R, Ferrarelli F, Hill S, Tononi G (2004). The sleep slow oscillation as a traveling wave. The Journal of Neuroscience 24:6862 -- 6870. Melamed O, Barak O, Silberberg G, Markram H, Tsodyks M (2008). Slow oscil- lations in neural networks with facilitating synapses. Journal of computational neuroscience 25:308 -- 316. Moreno-Bote R, Rinzel J, Rubin N (2007). Noise-induced alternations in an at- tractor network model of perceptual bistability. J Neurophysiol 98:1125 -- 39. Nakao H, Arai K, Kawamura Y (2007). Noise-induced synchronization and clus- tering in ensembles of uncoupled limit-cycle oscillators. Physical review letters 98:184101. Renart A, Moreno-Bote R, Wang XJ, Parga N (2007). Mean-driven and fluctuation-driven persistent activity in recurrent networks. Neural computa- tion 19:1 -- 46. Sanchez-Vives MV, McCormick DA (2000). Cellular and network mechanisms of rhythmic recurrent activity in neocortex. Nat Neurosci 3:1027 -- 34. Smeal RM, Ermentrout GB, White JA (2010). Phase-response curves and syn- chronized neural networks. Philosophical Transactions of the Royal Society B: Biological Sciences 365:2407 -- 2422. Steriade M, Nunez A, Amzica F (1993). A novel slow (¡ 1 hz) oscillation of neocortical neurons in vivo: depolarizing and hyperpolarizing components. The Journal of Neuroscience 13:3252 -- 3265. Tabak J, Senn W, O'Donovan MJ, Rinzel J (2000). Modeling of spontaneous activity in developing spinal cord using activity-dependent depression in an ex- citatory network. The Journal of Neuroscience 20:3041 -- 3056. Teramae Jn, Tanaka D (2004). Robustness of the noise-induced phase synchro- nization in a general class of limit cycle oscillators. Physical review letters 93:204103. Traub RD, Whittington MA, Stanford IM, Jefferys JG (1996). A mechanism for generation of long-range synchronous fast oscillations in the cortex . Vicente R, Gollo LL, Mirasso CR, Fischer I, Pipa G (2008). Dynamical relay- ing can yield zero time lag neuronal synchrony despite long conduction delays. Proceedings of the National Academy of Sciences 105:17157 -- 17162. Volgushev M, Chauvette S, Mukovski M, Timofeev I (2006). Precise long-range synchronization of activity and silence in neocortical neurons during slow-wave sleep. The Journal of Neuroscience 26:5665 -- 5672. Wang XJ (2001). Synaptic reverberation underlying mnemonic persistent activity. Trends Neurosci 24:455 -- 63.
1809.05196
1
1809
2018-09-13T22:27:24
Statistical Data Assimilation: Formulation and Examples from Neurobiology
[ "q-bio.NC", "physics.bio-ph" ]
For the Research Topic Data Assimilation and Control: Theory and Applications in Life Sciences we first review the formulation of statistical data assimilation (SDA) and discuss algorithms for exploring variational approximations to the conditional expected values of biophysical aspects of functional neural circuits. Then we report on the application of SDA to (1) the exploration of properties of individual neurons in the HVC nucleus of the avian song system, and (2) characterizing individual neurons formulated as very large scale integration (VLSI) analog circuits with a goal of building functional, biophysically realistic, VLSI representations of functional nervous systems. Networks of neurons pose a substantially greater challenge, and we comment on formulating experiments to probe the properties, especially the functional connectivity, in song command circuits within HVC.
q-bio.NC
q-bio
Statistical Data Assimilation: Formulation and Examples from Neurobiology Anna Miller 1,∗, Dawei Li 1 Jason Platt 1 Arij Daou 4 Daniel Margoliash 2 and Henry D. I. Abarbanel 1,3 1Department of Physics, University of California San Diego, La Jolla , CA, USA 2 Department of Anatomy and Organismal Biology, University of Chicago, Chicago , IL, USA 3 Marine Physical Laboratory (Scripps Institution of Oceanography) and Department of Physics, University of California, La Jolla, CA, USA 4 Department of Biomedical Engineering, American University of Beirut, Beirut, Lebanon Correspondence*: Anna Miller [email protected] ABSTRACT For the Research Topic Data Assimilation and Control: Theory and Applications in Life Sciences we first review the formulation of statistical data assimilation (SDA) and discuss algorithms for exploring variational approximations to the conditional expected values of biophysical aspects of functional neural circuits. Then we report on the application of SDA to (1) the exploration of properties of individual neurons in the HVC nucleus of the avian song system, and (2) characterizing individual neurons formulated as very large scale integration (VLSI) analog circuits with a goal of building functional, biophysically realistic, VLSI representations of functional nervous systems. Networks of neurons pose a substantially greater challenge, and we comment on formulating experiments to probe the properties, especially the functional connectivity, in song command circuits within HVC. Keywords: Data Assimilation, Neuronal Dynamics, HVC, Ion Channel Properties, Variational Annealing, Neuromorphic, VLSI 1 INTRODUCTION A broad class of 'inverse' problems presents itself in many scientific and engineering inquiries. The overall question addressed by these is how to transfer information from laboratory and field observations to candidate models of the processes underlying those observations. The existence of large, information rich, well curated data sets from increasingly sophisticated observations on complicated nonlinear systems has set new challenges to the information transfer task. Assisting with this challenge are new substantial computational capabilities. Together they have provided an arena in which principled formulation of this information transfer along with algorithms to effect the transfer have come to play an essential role. This paper reports on some efforts to meet this class of challenge within neuroscience. Many of the ideas are applicable much more broadly than our focus, and we hope the reader will find this helpful in their own inquiries. 1 Miller et al. In this special issue entitled Data Assimilation and Control: Theory and Applications in Life Sciences, of the journal Frontiers in Applied Mathematics and Statistics -- Dynamical Systems, we participate in the broader quantitative setting for this information transfer. The procedures are called "data assimilation" following its use in the effort to develop realistic numerical weather prediction models [1, 2] over many decades. We prefer the term 'statistical data assimilation' (SDA) to emphasize that key ingredients in the procedures involved in the transfer rest on noisy data and on recognizing errors in the models to which information in the noisy data is to be transferred. This article begins with a formulation of SDA with some additional clarity beyond the discussion in [3]. We also discuss some algorithms helpful for implementing the information transfer, testing model compatibility with the available data, and quantitatively identifying how much information in the data can be represented in the model selected by the SDA user. Using SDA will also remind us that data assimilation efforts are well cast as problems in statistical physics [4]. After the discussion of SDA, we turn to some working ideas on how to perform the high dimensional integrals involved in SDA. In particular we address the 'standard model' of SDA where data is contaminated by Gaussian noise and model errors are represented by Gaussian noise, though the integrals to be performed are, of course, not Gaussian. The topics include the approximation of Laplace [5, 6] and Monte Carlo methods. With these tools in hand, we turn to neurobiological questions that arise in the analysis of individual neurons and, in planning, for network components of the avian song production pathway. These questions are nicely formulated in the general framework, and we dwell on specifics of SDA in a realistic biological context. The penultimate topic we address is the use of SDA to calibrate VLSI analog chips designed and built as components of a developing instantiation of the full songbird song command network, called HVC. Lastly, we discuss the potential utlization of SDA for exploring biological networks. At the outset of this article we may expect that our readers from Physics and Applied Mathematics along with our readers from Neurobiology may find the conjunction of the two "strange bedfellows" to be incongruous. For the opportunity to illuminate the natural melding of the facets of both kinds of questions, we thank the editors of this special issue. 2 MATERIALS AND METHODS 2.1 General Overview of Data Assimilation We will provide a structure within which we will frame our discussion of transfer of information from data to a model of the underlying processes producing the data. We start with a observation window in time [t0, tF ] within which we make a set of measurements at times t = {τ1, τ2, ..., τk, ..., τF}; t0 ≤ τk ≤ tF . At each of these measurement times, we observe L quantities y(τk) = {y1(τk), y2(τk), ..., yL(τk)}. The number L of observations at each measurement time τk is typically less, often much less, than the number of degrees of freedom D in the observed system; D (cid:29) L. These are views into the dynamical processes of a system we wish to characterize. The quantitative characterization is through a model we choose. It describes the interactions among the states of the observed system. If we are observing the time course of a neuron, for example, we might measure the membrane voltage y1(τk) = Vm(τk) and the intracellular Ca2+ concentration y2(τk) = [Ca2+](τk). From these data This is a provisional file, not the final typeset article 2 Miller et al. we want to estimate the unmeasured states of the model as a function of time as well as estimate biophysical parameters in the model. The processes characterizing the state of the system (neuron) we call xa(t); a = 1, 2, ..., D ≥ L, and they are selected by the user to describe the dynamical behavior of the observations through a set of equations in continuous time = Fa(x(t), q), dxa(t) dt or in discrete time tn = t0 + n∆t; n = 0, 1, ..., N ; tN = tF via xa(tn+1) = xa(n + 1) = fa(x(tn), q) = fa(x(n), q), (1) (2) where q is a set of fixed parameters associated with the model. f (x(n), q) is related to F(x(t), q) through the choice the user makes for solving the continuous time flow for x(t) through a numerical solution method of choice [7]. Considering neuronal activity, Eq. (1) could be coupled Hodgkin-Huxley (HH) equations [8, 9] for voltage, ion channel gating variables, constituent concentrations, and other ingredients. If the neuron is isolated in vitro, such as by using drugs to block synaptic transmission, then there would be no synaptic input to the cell to describe. While if it is coupled to a network of neurons, their functional connectivity would be described in F(x(t), q) or f (x(n), q). Typical parameters might be maximal conductances of the ion channels, reversal potentials, and other time-independent numbers describing the kinetics of the gating variables. In many experiments L is only 1, namely, the voltage across the cell membrane, while D may be on the order of 100; Hence D (cid:29) L. As we proceed from the initiation of the observation window at t0 we must move our model equation variables x(0), Eq. (2), from t0 to τ1 where a measurement is made. Then using the model dynamics we move along to τ2 and so forth until we reach the last measurement time τF and finally move the model from x(τF ) to x(tF ). In each stepping of the model equations Eq. (2) we may make many steps of ∆t in time to achieve accuracy in the representation of the model dynamics. The full set of times tn at which we evaluate the model x(tn) we collect into the path of the state of the model through D-dimensional space: X = {x(0), x(1), ..., x(n), ..., x(N ) = x(F )}. The dimension of the path is (N + 1)D + Nq, where Nq is the number of parameters q in our model. It is worth a pause here to note that we have now collected two of the needed three ingredients to effect our transfer of the information in the collection of all measurements Y = {y(τ1), y(τ2), ..., y(τF )} to the model f (x(n), q) along the path X through the observation window [t0, tF ]: (1) data Y and (2) a model of the processes in Y, devised by our experience and knowledge of those processes. The notation and a visual presentation of this is found in Fig. (1). The third ingredient, comprised of methods to generate the transfer from Y to properties of the model, will command our attention throughout most of this paper. If the transfer methods are successful and, according to some metric of success, we arrange matters so that at the measurement times τk, the L model variables x(t) associated with y(τk) are such that xl(τk) ≈ yl(τk), we are not finished. We have then only demonstrated that the model is consistent with the known data Y. We must use the model, completed by the estimates of q and the state of the model at tF , x(tF ), to predict forward for t > tF , and we should succeed in comparison with measurements for y(τr) for τr > tF . As the measure of success of predictions, we may use the same metric as utilized in the observation window. Frontiers 3 Miller et al. Figure 1. A visual representation of the window t0 ≤ t ≤ tF in time during which L-dimensional observations y(τk) are performed at observation times t = τk; k = 1, ..., F . This also shows times at which the D-dimensional model developed by the user x(n + 1) = f (x(n), q) is used to move forward from time n to time n + 1: tn = t0 + n∆t; n = 0, 1, ..., N. D ≥ L. The path of the model X = {x(0), x(1), ..., x(n), ..., x(N ) = x(F )} and the collection Y of L-dimensional observations at each observation time τk, Y = {y(τ1), y(τ2), ..., y(τk), ..., y(τF} (y = {y1, y2, ..., yL}) is also indicated. As a small aside, the same overall setup applies to supervised machine learning networks [10] where the observation window is called the training set; the prediction window is called the test set, and prediction is called generalization. 2.1.1 The Data are Noisy; the Model has Errors Inevitably, the data we collect is noisy, and equally the model we select to describe the production of those data has errors. This means we must, at the outset, address a conditional probability distribution P (XY) as our goal in the data assimilation transfer from Y to the model. In [3] we describe how to use the Markov nature of the model x(n) → x(n + 1) = f (x(n), q) and the definition of conditional probabilities to derive the recursion relation: P (X(n + 1)Y(n + 1)) = P (y(n + 1), x(n + 1), X(n)Y(n)) P (y(n + 1)Y(n)) P (x(n + 1), X(n + 1)Y(n)) P (x(n + 1)x(n))P (X(n)Y(n)) = exp[CM I(y(n + 1), x(n + 1), X(n)Y(n))] • P (x(n + 1)x(n))P (X(n)Y(n)), This is a provisional file, not the final typeset article • (3) 4 Miller et al. (cid:104) (P (a,bc) (cid:105) where we have identified CM I(a, bc) = log . This is Shannon's conditional mutual information [11] telling us how many bits (for log2) we know about a when observing b conditioned on c. For us a = {y(n + 1)}, b = {x(n + 1), X(n + 1)}, c = {Y(n)}. P (ac) P (ac) Using this recursion relation to move backwards through the observation window from tF = t0 + N ∆t through the measurements at times τk to the start of the window at t0, we may write, up to factors independent of X P (XY) = P (y(τk)X(τk)) P (x(n + 1)x(n)) P (x(0)). (4) k=1 n=0 If we now write P (XY) ∝ exp[−A(X)] where A(X), the negative of the log likelihood, we call the action, then conditional expected values for functions along the path X are defined by (cid:26) F(cid:89) N−1(cid:89) E[G(X)Y] = (cid:104)G(X)(cid:105) = , (5) (cid:82) dX G(X)e−A(X) (cid:82) dX e−A(X) dX =(cid:81)N convenient expression (cid:26) F(cid:88) N−1(cid:88) (cid:27) (cid:27) n=0 dDx(n), and all factors in the action independent of X cancel out here. The action takes the A(X) = − log[P (y(τk)X(τk)] + log[P (x(n + 1)x(n))] − log[P (x(0))], (6) k=1 n=0 which is the sum of the terms which modify the conditional probability distribution when an observation is made at t = τk and the sum of the stochastic version of x(n) → x(n + 1) − f (x(n), q) and finally the distribution when the observation window opens at t0. What quantities G(X) are of interest? One natural one is the path G(X) = Xµ; µ = {a, n} itself; another is the covariance around that mean (cid:104)Xµ(cid:105) = ¯Xµ = (cid:104)Xµ(cid:105) : (cid:104)(Xµ − ¯Xµ)(Xν − ¯Xν)(cid:105). Other moments are of interest, of course. If one has an anticipated form for the distribution at large X, then G(X) may be chosen as a parametrized version of that form and those parameters determined near the maximum of P (XY). The action simplifies to what we call the 'standard model' of data assimilation when (1) observations y are related to their model counterparts via Gaussian noise with zero mean and diagonal precision matrix Rm, and (2) model errors are associated with Gaussian errors of mean zero and diagonal precision matrix Rf : A(X) = Rm(k) 2 (xl(τk) − yl(τk))2 + Rf (a) 2 (xa(n + 1) − fa(x(n), q))2. (7) If we have knowledge of the distribution P (x(0)) at t0 we may add it to this action. If we have no knowledge of P (x(0)), we may take its distribution to be uniform over the dynamic range of the model variables, then it, as here, is absent, canceling numerator and denominator in Eq. (5). Our challenge is to perform integrals such as Eq. (5). One should anticipate that the dominant contribution to the expected value comes from the maxima of P (XY) or, equivalently the minima of A(X). At such Frontiers 5 N−1(cid:88) D(cid:88) n=0 a=1 F(cid:88) L(cid:88) k=1 l=1 Miller et al. minima, the two contributions to the action, the measurement error and the model error, balance each other to accomplish the explicit transfer of information from the former to the latter. We note, as before, that when f (x(n), q) is nonlinear in X, as it always is in interesting examples, the expected value integral Eq. (5) is not Gaussian. So, some thinking is in order to approximate this high dimensional integral. We turn to that now. After consideration of methods to do the integral, we will return to a variety of examples taken from neuroscience. The two generally useful methods available for evaluating this kind of high dimensional integral are Laplace's method [5, 6] and Monte Carlo techniques [7, 12, 13]. We address them in order. We also add our own new and useful versions of the methods. 2.1.2 Laplace's Method To locate the minima of the action A(X) = − log[P (XY)] we must seek paths X(j); j = 0, 1, ... such that ∂A(X)/∂XX(j) = 0, and then check that the second derivative at X(j), the Hessian, is a positive definite matrix in path coordinates. The vanishing of the derivative is a necessary condition. Laplace's method [5] expands the action around the X(j) seeking the path X(0) with the smallest value of A(X). The contribution of X(0) to the integral Eq. (5) is approximately exp[A(X(1)) − A(X(0))] bigger than that of the path with the next smallest action. This sounds more or less straightforward; however, finding the global minimum of a nonlinear function such as A(X) is an NP-complete problem [14]. In a practical sense one cannot expect to succeed with such problems. However there is an attractive feature of the form of A(X) that permits us to accomplish more. We now discuss two algorithmic approaches to implementing Laplace's method. 2.1.3 Precision Annealing for Laplace's Method Looking at Eq. (7) we see that if the precision of the model is zero, Rf = 0, the action is quadratic in the L measured variables xl(n) and independent of the remaining states. The global minimum of such an action comes with xl(τk) = yl(τk) and any choice for the remaining states and parameters. Choose the path with these values of x(τk) and values from a uniform distribution of the other state variables and parameters covering the expected dynamic range of those, and call it path Xinit. In practice, we recognize that the global minimum of A(X) is degenerate at Rf = 0, so we select many initial paths. We choose NI of them, and initialize whatever numerical optimization program we have selected, to run on each of them. We continue to call the collection of NI paths Xinit. • Now we increase Rf from Rf = 0 to a small value Rf 0. Use each of the NI paths in Xinit as initial conditions for our numerical optimization program chosen to find the minima of A(X), and we arrive at NI paths X0. Evaluate A(X0) on all NI paths X0. • We increase Rf = Rf 0 → Rf 0α; α > 1, and now use the NI paths X0 as the initial conditions for our numerical optimization program chosen to find the minima of A(X), we arrive at NI paths X1. Evaluate A(X1) on all NI paths X1. • We increase Rf = Rf 0α → Rf 0α2. Now use the NI paths X1 as the initial conditions for our numerical optimization program chosen to find the minima of A(X), we arrive at NI paths X2. Evaluate A(X2) on all NI paths X2. • Continue in this manner increasing Rf to Rf = Rf 0αβ; β = 0, 1, ..., then using the selected numerical optimization program to arrive at NI paths Xβ. Evaluate A(Xβ) on all NI paths Xβ. This is a provisional file, not the final typeset article 6 Miller et al. • As a function of logα (cid:105) (cid:104) Rf Rf 0 display all NI values of A(Xβ) versus β for all β = 0, 1, 2, ...βmax. We call this method precision annealing (PA) [15, 16, 17, 18]. It slowly turns up the precision of the model collecting paths at each Rf that emerged from the degenerate global minimum at Rf = 0. In practice it is able to track NI possible minima of A(X) at each Rf . When not enough information is presented to the model, that is L is too small, there are many local minima at all Rf . This is a manifestation of the NP-completeness of the minimization of A(X) problem. None of the minima may dominate the expected value integral of interest. As L increases, and enough information is transmitted to the model, for large Rf one minimum appears to stand out as the global minimum, and the paths associated with that smallest minimum yields good √ predictions. We note that there are always paths, not just a single path, as we have a distribution of paths, NI of which are sampled in the PA procedure, within a variation of size 1/ Rm. A clear example of this is seen in [19] in a small, illustrative model. 2.1.4 "Nudging" within Laplace's Method In meteorology one approach to data assimilation is to add a term to the deterministic dynamics which move the state of a model towards the observations [20] xa(n + 1) = fa(x(n), q) + u(n)(yl(n) − xl(n))δal, (8) where u(n) > 0 and vanishes except where a measurement is available. This is referred to as 'nudging'. It appears in an ad hoc, but quite useful, manner. Within the structure we have developed, one may see that the 'nudging term' arises through the balance between the measurement error term and the model error term in the action. This is easy to see when we look at the continuous time version of the data assimilation standard model A(x(t), x(t)) = dt Rm(t, l) (xl(t) − yl(t))2 (cid:90) tF (cid:26) L(cid:88) t0 l=1 2 (cid:27) Rf (a) 2 ( xa(t) − Fa(x(t), q))2 . (9) D(cid:88) a=1 + (cid:20) The extremum of this action is given by the Euler-Lagrange equations for the variational problem [21] (cid:21)(cid:20) (cid:21) δab d dt + = Rm(a, t) Rf (a) xb(t) − Fb(x(t)) ∂Fb(x(t) ∂xa(t) δal(xl(t) − yl(t)), (10) in which the right hand side is the 'nudging' term appearing in a natural manner. Approximating the operator δab ∂xa(t) we can rewrite this Euler-Lagrange equation in 'nudging' form d dt + ∂Fb(x(t) Frontiers dxa(t) dt = Fa(x(t)) + u(t)δal(xl(t) − yl(t)). (11) 7 Miller et al. We will use both the full variation of the action, in discrete time, as well as its nudging form in our examples below. 2.1.5 Monte Carlo Methods Monte Carlo methods [22, 12, 18, 7] are well covered in the literature. We have not used them in the examples in this paper. However, the development of a precision annealing version of Monte Carlo techniques promises to address the difficulties with large matrices for the Jacobian and Hessians required in variational principles [23]. When one comes to network problems, about which we comment later, this method may be essential. 3 RESULTS 3.1 Using SDA to Analyze the Avian Song System We take our examples of the use of SDA in neurobiology from experiments on the avian song system. These have been performed in the University of Chicago laboratory of Daniel Margoliash, and we do not plan to describe in any detail the experiments nor the avian song production pathways in the avian brain. We give the essentials of the experiments and direct the reader to our references to develop the full biologically oriented idea why this system is enormously interesting. Essentially, however, the manner in which songbirds learn and produce their functional vocalization -- song -- is an elegant non-human example of a behavior that is cultural: the song is determined both by a genetic substrate and, interestingly, by refinement on top of that substrate by juveniles learning the song from their (male) elders. The analogs to learning speech in humans [24] are striking. Our avian singer is a zebra finch. They, as do most other songbirds, learn vocal expression through auditory feedback [25, 26, 27, 28, 24]. This makes the study of the song system a good model for learning complex behavior [29, 27, 30]. Parts of the song system are analogous to the mammalian basal ganglia and regions of the frontal cortex [27, 31, 32]. Zebra finch in particular have the attractive property of singing only a single learned song, and with high precision, throughout their adult life. Beyond the auditory pathways themselves, two neural pathways are principally responsible for song acquisition and production in zebra finch. The first is the Anterior Forebrain Pathway (AFP) which modulates learning. The second is a posterior pathway responsible for directing song production: the Song Motor Pathway (SMP) [26, 28, 33]. The HVC nucleus in the avian brain uniquely contributes to both of these [28]. There are two principal classes of projection neurons which extend from HVC: neurons which project to the robust nucleus of the arcopallium (HVCRA), and neurons which project to Area X (HVCX). HVCRA neurons extend to the SMP pathway and HVCX neurons extend to the AFP [28, 34]. These two classes of projection neurons combined with classes of HVC interneurons, make up the three broad classes of neurons within HVC. Fig. (2) [35] displays these structures in the avian brain. In vitro observations of each HVC cell type have been obtained through patch-clamp techniques making intracellular voltage measurements in a reduced, brain slice preparation [25]. In this configuration, the electrode can simultaneously inject current into the neuron while measuring the whole cell voltage response [36]. From these data, one can establish the physical parameters of the system [25]. Traditionally this is done using neurochemicals to block selected ion channels and measuring the response properties of others [37]. Single current behavior is recorded and parameters are found through mathematical fits of the This is a provisional file, not the final typeset article 8 Miller et al. data. This procedure has its drawbacks, of course. There are various technical problems with the choice of channel blockers. Many of even the modern channel blockers are not subtype specific [38] and may only partially block channels [39]. A deeper conceptual problem is that is difficult to know what channels one may have missed altogether. Perhaps there are channels which express themselves only outside the bounds of the experimental conditions. Our solution to such problems is the utilization of statistical data assimilation (SDA). This a method developed by meteorologists and others as computational models of increasingly large dynamical systems have been desired. Data assimilation has been described in our earlier sections. In this paper, we focus on the song learning pathway, reporting on experiments involving the HVCX neuron. The methods are generally applicable to the other neurons in HVC, and actually, to neurons seen as dynamical systems in general. We start with a discussion about the neuron model. First we demonstrate the utility of our precision annealing methods through the use of twin experiments. These are numerical experiments in which 'data' is generated through a known model (of HVCX), then analyzed via precision annealing. In a twin experiment, we know everything, so we can verify the SDA method by looking at predictions after a observation window in which the model is trained, and we may also compare the estimations of unobserved state variables and parameters to the ingredients and output of the model. Twin experiments are meant to mirror the circumstances of the real experiment. We start by taking the model that we think describes our physical system. Initial points for the state variables and parameters are chosen, which are used along with the model to numerically integrate forward in time. This leaves us with complete information about the system. Noise is added to a subset of the state variables to emulate the data to be collected in a lab experiment. We then perform PA on these simulated data, as if they were real data. The results of these numerical experiments can be used to inform laboratory experiments, and indeed help design them, by identifying the necessary measurements and stimulus needed to accurately electrophysiologically characterize a neuron. The second set of SDA analyses we report on using 'nudging', as described above, to estimate some key biophysical properties of HVCX neurons from laboratory data. This SDA procedure is applied to HVCX neurons in two different birds. The results show that though each bird is capable of normal vocalization, their complement of ion channel strengths is apparently different. We report on a suggestive example of this, leaving a full discussion to [40]. In order to obtain good estimation results, we must choose a forcing or stimulus with the model in mind: the dynamical range of the neuron must be thoroughly explored. This suggests a few key properties of the stimulus: • The current waveform of Iinjected(t) must have sufficient amplitudes (±) and must be applied sufficiently long in time that it explores the full range of the neuron variation. • The frequency content of the stimulus current must be a low enough that it does not exceed the low-pass cutoff frequency associated with the RC time constant of the neuron. This cutoff is typically in the neighborhood of 50-100Hz. • The current must explore all time scales expressed in the neuron's behavior. Frontiers 9 Miller et al. 3.2 Analysis of HVCX Data The model for an HVCX neuron is substantially taken from [25] and described in our Appendix. We now use this model in a 'twin experiment' in which PA is utilized, and then using 'nudging' we present the analysis of experimental data on two Zebra Finch. 3.2.1 Twin Experiment on HVCX Neuron Model A twin experiment is a synthetic numerical experiment meant to mirror the conditions of a laboratory experiment. We use our mathematical model with some informed parameter choices in order to generate numerical data. Noise is added to observable variables in the model, here V (t). These data are then put through our SDA procedure to estimate parameters and unobserved states of the model. The neuron model is now calibrated or completed. Using the parameters and the full state of the model at the end tF of an observation window [t0, tF ], we take a current waveform Iinjected(t ≥ tF ) to drive the model neuron and predict the time course of all dynamical variables in the prediction window [tF , ...]. This validation of the model is the critical test of our SDA procedure, here PA. In a laboratory experiment we have no specific knowledge of the parameters in the model and, by definition, cannot observe the unobserved state variables; here we can do that. So, 'fitting' the observed data within the observation window [t0, tF ] is not enough, we must reproduce all states for t ≥ tF to test our SDA methods. We use the model laid out in the Appendix. We assume that the neuron has a resting potential of −70 mV and set the initial values for the voltage and each gating variable accordingly. We assume that the internal calcium concentration of the cell is Cin = 0.1 µM. We use an integration time step of 0.02 ms and Figure 2. A drawing of the Song Production Pathway and the Anterior Forebrain Pathway of avian songbirds. Parts of the auditory pathways are shown in grey. Pathways from the brainstem that ultimately return to HVC are not shown. The HVC network image is taken from [35]. Reprinted by permission from Copyright Clearance Center: Springer Nature, Nature Reviews Neuroscience, Auditory Feedback in Learning and Maintenance of Vocal Behavior, M. S. Brainard and A. J. Doupe, 2000. This is a provisional file, not the final typeset article 10 Miller et al. integrate forward in time using an adaptive Runge-Kutta method [7]. Noise is added to the voltage time course by sampling from a Gaussian distribution N (0, 2) in units of mV. Figure 3. Stimulating current Iinjected(t) presented to the HVCX model. The waveform of the injected current was chosen to have three key attributes: (1) It is strong enough to cause spiking in the neuron, (2) it dwells a long time in a hyperpolarizing region, and (3) its overall frequency content is low enough to not be filtered out by the neuron's RC time constant. On this last point, a neuron behaves like an RC circuit, it has a cut off frequency limited by the time constant of the system. Any input current which has a frequency higher than that of the cut off frequency won't be 'seen' by the neuron. The time constant is taken to be the time it takes to spike and return back to 37% above its resting voltage. We chose a current where the majority of the power spectral density exists below 50 Hz. The first two seconds of our chosen current waveform is a varying hyperpolarizing current. In order to characterize Ih(t) and ICaT (t), it is necessary to thoroughly explore the region where the current is active. Ih(t) is only activated when the neuron is hyperpolarized. The activation of Ih(t) deinactivates ICaT (t), thereby allowing its dynamics to be explored. In order to characterize IN a(t) and IK(t), it is necessary to cause spiking in the neuron. The depolarizing current must be strong enough to hit the threshold potential for spike activation. The parameters used to generate the data used in the twin experiment are in Table (1), and the injection current data and the membrane voltage response may be seen in figures Fig. (3) and Fig. (4). The numbers chosen for the data assimilation procedure in this paper are α = 1.4 and β ranging from 1 to 100. Rf,0,V = 10−4 for voltage and Rf,0,j = 1 for all gating variables. These numbers are chosen Frontiers 11 Miller et al. Figure 4. Response of the HVCX model membrane voltage to the selected Iinjected(t). The displayed time course V (t) has no added noise. because the voltage range is 100 times large than the gating variable range. Choosing a single Rf,0 value would result in the gating variable equations being less enforced than the voltage equation by a factor of 104. The α and β numbers are chosen because we seek to make Rf sufficiently large. The α and β values Rf 0 chosen allow Rf Rf 0 to reach 1015. During estimation we instructed our methods to estimate the inverse capacitance and estimate the ratio g(cid:48) = g instead of g and Cm independently. That separation can present a challenge to numerical procedures. We also estimated the reversal potentials as a check on the SDA method. Cm Within our computational capability we can reasonably perform estimates on 50,000 data points. This captures a second of data when ∆t = 0.02 ms. However, there are time constants in the model neuron which are on order 1 second. In order for us to estimate the behavior of these parameters accurately, we need to see multiple instances of the full response. We need a window on the order of 2-3 seconds. We can obtain this by downsampling the data. We know from previous results that downsampling can lead to better estimations [41]. We take every ith data point, depending on the level of down sampling we want to do. In this data assimilation run, we downsampled by a factor of 4 to incorporate 4 seconds of data in the estimation window. We look at a plot of the action as a function of β; that is, log[Rf /Rf 0]. We expect to see a leveling of of the action [17] as a function of Rf . If the action becomes independent of Rf , we can then explore how well our parameter estimations perform when integrating them forward as predictions of the calibrated This is a provisional file, not the final typeset article 12 Miller et al. Figure 5. Action Levels of the standard model action for the HVCX neuron model discussed in the text. We see that the action rises to a 'plateau' where it becomes quite independent of Rf . The calculation of the action uses PA with α = 1.4 and Rf 0 = Rm. NI = 100 initial choices for the path Xinit were used in this calculation. For small Rf one can see the slight differences in action level associated with local minima of A(X). model. Looking at the action plot in Fig. (5), we can see there is a region in which the action appears to level off, around β = 40. It is in this region where we look for our parameter estimates. We examine all solutions around this region of β and utilize their parameter estimates in our predictions. We compare our numerical prediction to the "real" data from our synthetic experiment. We evaluate good predictions by finding the correlation coefficient between these two curves. This metric is chosen instead of a simple root mean square error because slight variations in spike timings yield a high amount of error even if the general spiking pattern is correct. The prediction plot and parameters for the best prediction can be seen in Fig. (6) and Table (2). The voltage trace in red is the estimated voltage after data assimilation is completed. It is overlayed on the synthetic input data in black. The blue time course is a prediction, starting at the last time point of the red estimated V (t) trace and using the parameter estimates for t ≤ 4000 ms. The red curve matches the computed voltage trace quite well. There is no deviation in the frequency of spikes, spike amplitudes, or the hyperpolarized region of the cell. Looking at the prediction window, we can see that there is some deviation in the hyperpolarized voltage trace around 9000 ms. This is an indication that our parameter estimates for currents activated in this region are not correct. Comparing parameters, we can see that Eh is estimated as lower than its actual value. Despite that, we still are able to reproduce neuron behavior fairly well. Frontiers 13 Miller et al. Figure 6. Results of the 'twin experiment' using the model HVCX neuron described in the Appendix. Noise was added to data developed by solving the dynamical equations. The noisy V (t) was presented to the precision annealing SDA calculation along with the Iinjected(t) in the observation window t0 = 0ms, tF = 4000 ms. The noisy model voltage data is shown in black, and the estimated voltage is shown in red. For t ≥ 4000ms we show the predicted membrane voltage, in blue, generated by solving the HVCX model equations using the parameters estimated during SDA within the observation window. 3.2.2 Analysis of Biophysical Parameters from HVCX Neurons in two Zebra Finch Our next use of SDA employs the 'nudging' method described in Eq. (8). In this section we used some of the data [40] taken in experiments on multiple HVCX neurons from different zebra finches. The questions we asked was whether we could, using SDA, identify differences in biophysical characteristics of the birds. This question is motivated by prior biological observations [40]. Using the same HVCX model as before, we estimated the maximal conductances {gN a, gK, gCaT , gSK, gh} holding fixed other kinetic parameters and the Nernst/Reversal potentials. The baseline characteristics of an ion channel are set by the properties of the cell membrane and the complex proteins penetrating the membrane forming the physical channel. Differences among birds would then come from the density or numbers of various channels as characterized by the maximal conductances. If such differences were identified, it would promote further investigation of the biologically exciting proposition that these differences arise in relation to some aspect of the song learning experience of the birds [40]. In Fig. (7) and Fig. (8) we display the stimulating current and membrane voltage response from one of 9 neurons in our large sample. The analysis using SDA was of four neurons from one bird and seven neurons from another. The results for {gCaT , gN a, gSK} is displayed in Fig. (9). The maximal conductances from This is a provisional file, not the final typeset article 14 Miller et al. one bird are shown in blue and from the other bird, in red. There is a striking difference between the distributions of maximal conductances. We do not propose here to delve into the biological implications of these results. Nevertheless, we note that the neurons from each bird occupy a small but distinct region of the parameter space (Fig. 9). This result and its implications for birdsong learning, and more broadly for neuroscience, are described in [40]. Here, however, we display this result as an example of the power of SDA to address a biologically important question in a systematic, principled manner beyond what is normally achieved in analyses of such data. To fully embrace the utility of SDA for these experiments, however, further work is needed. A limitation of the present result is that the SDA estimates for gSK for a subset of the neurons/observations for Bird One reach the bounds of the observation window (Fig. 9). Addressing such issues would be prelude to the exciting possibility of estimating more parameters than just the principle ion currents in the Hodgkin- Huxley equations. This could use SDA numerical techniques to calculate, over hours or days, estimates of parameters that could require months or years of work to measure with traditional biological and biophysical approaches, in some cases requiring specialized equipment beyond that available for most in vitro recording set ups. In contrast, applying SDA to such data sets requires only a computer. Figure 7. One of the library of stimuli used in exciting voltage response activity in an HVCX neuron. The cell was prepared in vitro, and a single patch clamp electrode injected Iinjected(t) (this waveform) and recorded the membrane potential. Frontiers 15 05001000150020002500300035004000-150-100-50050100150200Iapplied(t) (pA)Time (Units of 0.02 ms) Stimulating Current #7 Miller et al. Figure 8. The voltage response from Iapplied, Fig. (7). One of the library of stimuli used in exciting voltage response activity in an HVCX neuron. The cell was prepared in vitro, and a single patch clamp electrode injected Iinjected(t) (this waveform) and recorded the membrane potential. 3.3 Analysis of Neuromorphic VLSI Instantiations of Neurons An ambitious effort in neuroscience is the creation of low power consumption analog neural-emulating VLSI circuitry. The goals for this effort range from the challenge itself to the development of fast, reconfigurable circuitry on which to incorporate information revealed in biological experiments for use in • creating model neural circuits of 'healthy' performance to be compared to subsequent observations on the same circuitry informed by 'unhealthy' performances. If the comparison can be made rapidly and accurately, the actual instantiations in the VLSI circuitry could be used to diagnose the changes in neuron properties and circuit connectivity perhaps leading to directions for cures, and • in creating VLSI realizations of neural circuits with desired functions -- say, learning syntax in interesting sequences -- might allow those functions to be performed at many times increased speed than seen in the biological manifestation. If those functionalities are of engineering utility, the speed up could be critical in applications. One of the curious roadblocks in achieving critical steps of these goals is that after the circuitry is designed and manufactured into VLSI chips, what comes back from a fabrication plant is not precisely what we designed. This is due to the realities of the manufacturing processes and not inadequacies of the designers. This is a provisional file, not the final typeset article 16 0500100015002000250030003500-100-80-60-40-2002040HVCX Voltage Response Stimulus 7V(t) (mV)Time (Units of 0.02 ms) Miller et al. Figure 9. A three dimensional plot of three of the maximal conductances estimated from HVCX cells using the stimulating current shown in Fig. (7). Membrane voltage responses from five neurons from one bird were recorded many times, and membrane voltage responses from four neurons from a second bird were recorded many times. One set of maximal conductances {gN a, gCaT , gSK} are shown. The estimates from Bird 1 are in red-like colors, and the estimates from Bird 2 are in blue-like colors. This is just one out of a large number of examples discussed in detail in [40]. To overcome this barrier in using the VLSI devices in networks, we need an algorithmic tool to determine just what did return from the factory, so we know how the nodes of a silicon network are constituted. As each chip is an electronic device built on a model design, and the flaws in manufacuring are imperfections in the realization of design parameters, we can use data from the actual chip and SDA to estimate the actual parameters on the chip. SDA has an advantageous position here. If we present to the chip input signals with much the same design as we prepared for the neruobiological experimets discussed in the previous section, we can measure everything about each output from the chip and use SDA to estimate the actual parameters produced in manufacturing. Of course, we do not know those paramters a priori so after estimating the parameters, thus 'calibrating' the chip, we must use those estimated parameters to predict the response of the chip to a new stimulating currents. That will validate (or not) the completion of the model of the actual circuitry on the chip and permit confidence in using it in building interesting networks. We have done this on chips produced in the laboratory of Gert Cauwenberghs at UCSD using PA [42, 41] and using 'nudging' as we now report. The chip we worked with was designed to produce the simplest spiking neuron, namely one having just Na, K, and leak channels [8, 9] as in the original HH experiments. This neuron has four state variables Frontiers 17 0.00.20.40.60.81.01.21.41.61.82.03004005006007002468101214gNagSKgCaT Bird One Bird Two Miller et al. {V (t), m(t), h(t), n(t)}: dt = gN am3(t)h(t)(EN a − V (t)) + gKn4(t)(EK − V (t)) dV (t) C +gL(EL − V (t)) + Iinjected(t) in which the gating variables w(t) = {m(t), h(t), n(t)} satisfy dw(t) (w∞(V (t)) − w(t)) = dt τw(V (t)) , (12) and the functions w∞(V ) are discussed in depth in [8, 9]. In our experiments on a 'NaKL' chip we used the stimulating current displayed in Fig. (10), Figure 10. This waveform for Iinjected(t) was used to drive the VLSI NaKL neuron after receipt from the fabrication facility. and measured all the neural responses {V (t), m(t), h(t), n(t)}. These observations were presented to the designed model within SDA to estimate the parameters in the model. We then tested/validated the estimations by using the calibrated model to predict how the VLSI chip would respond to a different injected current. In Fig. (11) we show the observed Vdata(t) in black, the estimation of the voltage through SDA in red, and the prediction of V (t) in blue for times after the end of the observation window. This is a provisional file, not the final typeset article 18 Miller et al. While one can be pleased with the outcome of these procedures, for our purposes we see that the use of our SDA algorithms gives the user substantial confidence in the functioning characteristics of the VLSI chips one wishes to use at the nodes of a large, perhaps even very large, realization of a desired neural circuit in VLSI. We are not unaware of the software developments to allow efficient calibration of very large numbers of manufactured silicon neurons. A possible worry about also determining the connectivity, both the links and their strength and time constants, may be alleviated by realizing these links through a high bandwidth bi-directional connection of the massive array of chips and the designation of connectivity characteristics on an off-chip computer. Figure 11. The NaKL VLSI neuron was driven by the waveform for Iinjected(t) seen in Fig. (10). The four state variables {V (t), m(t), h(t), n(t)} for the NaKL model were recorded and used in an SDA 'nudging' protocol to estimate the parameters of the model actually realized at the manufacturing facility. Here we display the membrane voltages: {Vdata(t), Vest(t), Vpred(t)} -- the observed membrane voltage response when Iinjected(t) was used, the estimated voltage response using SDA, and finally the predicted voltage response Vpred(t) from the calibrated model actually on the VLSI chip. In a laboratory experiment, only this attribute of a neuron would be observable. Part of the same analysis is the ability to observe, estimate and predict the experimentally unobservable gating variables. This serves, in this context, as a check on the SDA calculations. The Na inactivation variable h(t) is shown in Fig. (12) as its measured time course hdata(t) in black, its estimated time course hest(t) in red, and its predicted time course hpred(t) in blue. Frontiers 19 Miller et al. Figure 12. The NaKL VLSI neuron was driven by the waveform for Iinjected(t) seen in Fig. (10. The four state variables {V (t), m(t), h(t), n(t)} for the NaKL model were recorded and used in an SDA 'nudging' protocol to estimate the parameters of the model actually realized at the manufacturing facility. Here we display the Na inactivation variable h(t): {hdata(t), hest(t), hpred(t)} -- the observed h(t) time course when Iinjected(t) was used, the estimated h(t) time course using SDA, and finally the predicted h(t) time course from the calibrated model actually on the VLSI chip. In a laboratory experiment, this attribute of a neuron would be unobservable. Note we have rescaled the gating variable from its natural range 0 ≤ h(t) ≤ 1) to the range within the VLSI chip. The message of this Figure is in the very good accuracy and prediction of an experimentally unobservable time course. 4 DISCUSSION Our review of the general formulation of statistical data assimilation (SDA) started our remarks. Many details can be found in [3] and subsequent papers by the authors. Recognizing that the core problem is to perform, approximately of course, the integral in Eq. (5) is the essential take away message. This task requires well 'curated' data and a model of the processes producing the data. In the context of experiments in life sciences or, say, aquisition of data from earth system sensors, curation includes an assessment of errors and the properties of the instruments as well. One we have the data and a model, we still need a set of procedures to transfer the information from the data to the model, then test/validate the model on data not used to train the model. The techniques we covered are general. Their application to examples from neuroscience comprise the second part of this paper. In the second part we first address properties of the avian songbird song production pathway and a neural control pathway modulating the learning and production of functional vocalization -- song. We focus our This is a provisional file, not the final typeset article 20 Miller et al. attention on one class of neurons, HVCX, but have also demonstrated the utility of SDA to describe the response properties of other classes of neurons in HVC, such as HVCRA [43] and HVCI [44]. Indeed, the SDA approach is generally applicable wherever there is insight to relate biophysical properties of neurons to their dynamics through Hodgkin Huxely equations. Our SDA methods considered variational algorithms that seek the highest conditional probability distributions of the model states and parameters conditioned on the collection of observations over a measurement window in time. Other approaches, especially Monte Carlo algorithms were not discussed here, but are equally attractive. We discussed methods of testing models of HVCX neurons using "twin experiments" in which a model of the individual neuron produces synthetic data to which we add noise with a selected signal-to-noise-ratio. Some state variable time courses from the library of these model produced data, for us the voltage across the cell membrane, is then part of the action Eq. (7), specifically in the measurement error term. Errors in the model are represented in the model error term of the action. Using a precision annealing protocol to identify and track the global minimum of the action, the successful twin experiment gives us confidence in this SDA method from information trans from data to the model. We then introduced a 'nudging' method as an approximation to the Euler-Lagrange equations derived from the numerical optimization of the action Eq. (7) -- this is Laplace's method in our SDA context. The nudging method, introduced in meteorology some time ago, was used to distinguish between two different members of the Zebra Finch collection. We showed, in a quite preliminary manner, that the two, unrelated birds of the same species, express different HVC network properties as seen in a critical set of maximal conductances for the ion channels in their dynamics. Finally we turned to a consideration of the challenge of implementing in VLSI technology the neurons in HVC towards the goal of building a silicon-HVC network. The challenge at the design and fabrication stages of this effort where illuminated by our use of SDA to determine what was actually returned from the manufacturing process for our analog neurons. 4.1 Moving Forward to Network Analysis Finally, we have a few comments associated with the next stage of analysis of HVC. In this, and previous papers, we analyzed individual neurons in HVC. These analyses were assisted by our using SDA, through twin experiments, to design laboratory experiments though the selection of effective stimulilaing injected currents. Having characterized the electrophysiology of an individual neuron within the framework of Hodgkin- Huxley (HH) models, we may now proceed beyond the study of individual neurons [45] in vitro. Once we have characterized an HVC neuron through a biophysical HH model, we may then use it in vivo as a sensor of the activity of the HVC network where it is connected to HVCRA, HVCI, and other HVCX neurons. The schema for this kind of experiment is displayed in Fig. (13). These experiments require the capability to perform measurements on HVC neurons in the living bird. That capability is available, and experiments as suggested in our graphic are feasible, if challenging. The schematic indicates that the stimulating input to the experiments is auditory signals, chosen by the user, presented to the bird's ear and reaching HVC through the auditory pathway. The stimuli from this signal is then distributed in a manner to be deduced from experiment and then produces activity in the HVC network that we must model. The goal is, at least initially, to establish, again within the models we Frontiers 21 Miller et al. develop, the connectivity of HVC neuron classes as it manifests itself in the function of the network. We have some information about this [46, 47], and these results will guide the development of the HVC model used in these whole-network experiments. An important point to address is what changes to the in vitro model might be necessary to render it a model for in vivo activity. Figure 13. A cartoon-like idea of an experiment to probe the HVC network. In this graphic three neuron populations of {HV CX , HV CI , HV CRA} neurons are stimulated by auditory signals P (t) presented to the bird in vivo. This drives the auditory pathway from the ear to HVC, and the network activity is recorded from a calibrated, living HVCX neuron, used here as a sensor for network activity. While the experiment is now possible, the construction of a model HVC network will proceed in steps of complexity using simple then more biophysiclly realistic neuron models and connections among the nodes (neurons) of the network. From libraries of time courses of P (t), chosen by the user, and responses of V (t) in the 'sensor' HVCX neuron, we will use SDA to estimate properties of the network. This is a provisional file, not the final typeset article 22 Miller et al. APPENDIX The equations describing the HVCX neuron dynamics are taken from the work of Daou [25]. That paper also has an extensive account of his experiments on the other two major classes of neurons in HVC. The equations are of Hodgkin-Huxley (HH) form for a neuron without spatial extent; this is called a one- compartment model. It is meant to apply to neurons in isolation of the network, here HVC, in which they sit in vivo. The dynamical variables include the observable quantities: voltage across the cell membrane, V (t) and the intracellular concentration of [Ca2+]in(t) = C(t). V (t) is directly connected to action potentials or voltage spikes that communicate among cells in a network; the time scale of these spikes is a few ms. C(t) provides a slow background modulation that raises the cells potential (depolarizes the cell) or lowers it (hyperpolarizes the cell) on time scales as long as 10's of ms. The voltage equation, conservation of charge, relates the capacitance of the cell membrane Cm as it separates concentrations of ions within and without the cells to the various currents which contain the nonlinear voltage dependence of the permeability of ions to passing into and out of the cell. The model represents these ion currents: {N a, K, Ca} in several different ways. The general form of an HH current is (t)(Erev−ion − V (t)), ion ion (t)hinteger2 Iion(t) = gionminteger1 (13) where the reversal potential is the equilibrium Nernst potential [8, 9]. The gating variables {m(t), h(t)} lie between zero and one and represent the amount the ion channel is open relative to the maximum opening it may have. The maximal conductance gion represent the number or density of ion channels in the neuron model. This form of ion current applies when the concentrations of the ion are not significantly different outside and inside the cell. This is not so for Ca2+ ions, so there we use the Goldman-Hodgkin-Katz (GHK) form of the current [48]. (cid:32) (cid:33) Ia(t) = −Paz2 aF 2 [ion]in(t) − [ion]oute−zaF V (t)/RT 1 − e−zaF V (t)/RT , (14) for iona. za is the charge on the ion, F the Faraday constant, R the gas constant, T the temperature, and Pa the permeability of the cell membrane to ion a. We use an approximation to the GHK equations for the two types of Calcium currents selected in [25] (cid:18) (cid:19) [Ca]out e2F V (t)/RT − 1 (cid:18) ICaL(t) = gCaLV s2∞(V (t)) ICaT (t) = gCaT V (t)[aT ]3∞(V )[bT ]3∞(rA T ) (cid:17) − bT∞(rT ) = (cid:16) rT −θb 1 [Ca]out e2F V (t)/RT − 1 (cid:16)−θb 1 (cid:17) 1 + e σb 1 + e σb (cid:19) (15) aT and bT are instantaneous activating and inactivating gating variables, respectively. rT is a slow gating variable which takes the same functional form as aT and other gating variables m(t) and h(t). These gating Frontiers 23 Miller et al. variables w(t) satisfy a first order kinetic equation in which w∞(V ) = 1 2 1 − tanh dw(t) dt = w∞(V (t)) − w(t)) τw(V (t)) (cid:20) , (cid:18) V − θw (cid:19)(cid:21) (16) (17) , 2σw for all gating variables except h∞(V ) appearing in IN a(t) [25]. θw is the half-activation voltage and σw controls the slope of the activation function. For fast gating variables, such as m of IN a, and s of ICaL we replace the time dependence by W∞(V ). τw(V ) is the time constant of each gating variable. Time constants for the n and hp gating variables (these names refer to [25]) are given below, where ¯τw is an average time constant. Our model differs from [25] by one time constant. Instead, τrs(V ) takes the form presented here: τw(V ) = cosh for n or hp (cid:16) V −θw ¯τw 2σw (cid:17) (cid:18) (cid:19)(cid:19) (cid:18) V (t) + 80 −21 1 − tanh2 τrs(V ) = 0.1 + 193.0 τrf = prf −7.4(V +70) V +70−0.8 −1 e + 65e V +56−23 τrT (V ) = τr0 + 1 + e (cid:17) τr1 (cid:16) V −θrT σrT For our choice of ion currents we follow the results of experimental data [49, 50, 51] and generally reproduce the model listed in [25]. HVCX spiking properties include fast rectifying current, sag in response to hyperpolarizing current, and spike frequency adaption in response to depolarizing current. C dV (t) dt = IN a(t) + IK(t) + IL(t) + ICaT (t) + ICaL(t) +IA(t) + ISK(t) + Ih(t) + IN ap(t) + Iinjected(t) (18) IN a(t) and IK(t) are the standard HH currents. They produce fast spiking in response to injected currents. IL(t) is a leak current meant to capture all linear currents of the neuron. ICaT (t) is a low threshold T-type calcium current that causes rebound depolarization in cooperation with Ih(t). ICaL(t) is a high threshold L-type calcium current. ICaL(t) works in conjunction with ISK(t), a calcium concentration dependent potassium current, to create frequency adaptation in neuron spiking. IA(t) is an A-type potassium current. IN ap(t) is a persistent sodium current. From the model presented in [25], we eliminate IKN a(t), a sodium dependent potassium current, and rewrite all sigmoidal functions as hyperbolic tangents. The mass conservation equation for Ca2+ is written as dC(t) dt = (ICaT (t) + ICaL(t)) + kCa(bCa − C(t)), again following [25]. This is a provisional file, not the final typeset article (19) 24 Miller et al. TABLES Parameter gN a EN a gK EK gCaT gSK gH EH Cm θa σa θrs σrs θrT σrT τr0 kCa Parameter Value 450 nS gL 50 mV EL 50 nS gN ap -90 mV gCaL 2.65 nS θm 6 nS σm 4 nS θn -30 mV σn 100 pF ¯τn -20 mV θrf -10 mV σrf -105 mV θaR 25 mV σaT -67 mV θrrT 2 mV σrrT 200 ms τr1 0.3 ms−1 bCa Value Parameter 2 nS kf -70 mV θmp 1 nS σmp 19 nS θs -35 mV σs -5 mV θhp -30 mV σhp -5 mV ¯τhp 10 ms θe -105 mV σe 5 mV τe -65 mV θb -7.8 mV σb 68 mV f 2.2 mV  87.5 ms prf 0.1 µM ks Value 0.3 -40 mV -6 mV -20 mV -0.05 mV -48 mV 6 mV 1000 ms -60 mV 5 mV 20 ms 0.4 mV -0.1 mV 0.1 0.0015 µM pA·ms 100 0.5 µM Table 1. Parameter values used to numerically generate the HVCX data. The source of these values comes from [25]. Data was generated using an adaptive Runge-Kutta method, and can be seen in Fig. (3) and Fig. (4). Parameter g(cid:48) N a EN a g(cid:48) K EK g(cid:48) g(cid:48) CaT g(cid:48) h Eh Cinv SK Bounds 0.1, 10 1, 100 0.01, 5 -140, -10 0.001, 1 0.001, 1 0.001, 1 -100, -1 0.001, 0.5 4.5 50 0.5 -90 Best Estimate Actual Value Units nS/pF mV nS/pF mV nS/pF nS/pF nS/pF mV pF−1 4.98 43.2 0.907 -127.4 0.0326 0.0373 0.0432 -44.1 0.011 0.0265 0.06 0.04 -30 0.01 Table 2. Parameter Estimates from the Best Predictions. The best prediction is chosen by finding the highest correlation coefficient between the predicted voltage and "real" voltage. This comparison can be made on experimental data. It represents an attractive alternative to the familiar least squares metric commonly used. That metric is very sensitive to spike times in data with action potentials: small errors in spike times may result in large errors in a least squares metric. Frontiers 25 Miller et al. REFERENCES [1] Lorenz EN. Predictability: A problem partly solved. Palmer T, Hagedorn R, editors, Predictability of weather and climate (Cambridge) (2006). [2] Evensen G. Data Assimilation: The Ensemble Kalman Filter (Springer) (2009). [3] Abarbanel HDI. Predicting the Future: Completing Models of Observed Complex Systems (Springer) (2013). [4] Tong D. Statistical physics (2011). Available at www.damtp.cam.ac.uk/user/tong/statphys/sp.pdf. [5] Laplace PS. Memoir on the probability of causes of events. Math´ematique et de Physique,Tome Sixi´eme (1774) 621 -- 656. [6] Laplace P. Memoir of the probability of causes of events. Statistical Science 1 (1986) 365 -- 378. Translation to English by S. M. Stigler. [7] Press WH, Teukolsky SA, Vetterling WT, Flannery BP. Numerical Recipes: The Art of Scientific Computing, Third Edition (Cambridge University Press) (2007). [8] Johnston D, Wu SMS. Foundations of Cellular Neurophysiology (Bradford Books, MIT Press) (1995). [9] Sterratt D, Graham B, Gillies A, Willshaw D. Principles of Computational Modelling in Neuroscience (Cambridge University Press) (2011), 390 . [10] Abarbanel HDI, Rozdeba PJ, Shirman S. Machine learning: Deepest learning as statistical data assimilation problems. Neural Computation 30 (2018) 2025 -- 2055. [11] Fano RM. Transmission of Information; A Statistical Theory of Communication (MIT Press) (1961). [12] Kostuk M, Toth BA, Meliza CD, Margoliash D, Abarbanel HDI. Dynamical estimation of neuron and network properties ii: path integral monte carlo methods. Biological Cybernetics 106 (2012) 155 -- 167. [13] Neal R. Mcmc using hamiltonian dynamics. Gelman A, Jones G, Meng XL, editors, Handbook of Markov Chain Monte Carlo (Chapman and Hall; CRC Press), chap. 5 (2011). [14] Murty KG, Kabadi SN. Some np-complete problems in quadratic and nonlinear programming. Mathematical Programming 39 (1987) 117 -- 129. [15] Ye J, Kadakia N, Rozdeba PJ, Abarbanel HDI, Quinn JC. Improved variational methods in statistical data assimilation. Nonlinear Processes in Geophysics 22 (2015) 205 -- 213. doi:10.5194/ npg-22-205-2015. [16] Ye J, Rey D, Kadakia N, Eldridge M, Morone UI, Rozdeba P, et al. Systematic variational method for statistical nonlinear state and parameter estimation. Phys. Rev. E 92 (2015) 052901. doi:10.1103/ PhysevE.92.052901. [17] Ye J. Systematic Annealing Approach for Statistical Data Assimilation. Ph.D. thesis, University of California San Diego (2016). [18] Quinn JC. A path integral approach to data assimilation in stochastic nonlinear systems. Ph.D. thesis, University of California San Diego (2010). [19] Shirman S. Strategic Monte Carlo and Variational Methods in Statistical Data Assimilation for Nonlinear Dynamical Systems. Ph.D. thesis, University of California San Diego (2018). [20] Anthes R. Data assimilation and initialization of hurricane prediction models. J. Atmos. Sci. 31 (1974). [21] Gelfand IM, Fomin SV. Calculus of Variations (Dover Publications, Inc.) (1963). [22] Metropolis N, Rosenbluth AW, Rosenbluth MN, Teller AH, Teller E. Equation of State Calculations by Fast Computing Machines. J. Chem. Phys. 21 (1953) 1087 -- 1092. doi:10.1063/1.1699114. [23] Wong AS, Hao K, Abarbanel HDI. Precision annealing for monte carlo methods: Application to the shallow water equations (2018). In Preparation, 2018. This is a provisional file, not the final typeset article 26 Miller et al. [24] Doupe AJ, Kuhl PK. Birdsong and human speech: Common themes and mechanisms. Annual Review of Neuroscience 22 (1999) 567 -- 631. doi:10.1146/annurev.neuro.22.1.567. PMID: 10202549. [25] Daou A, Ross M, Johnson F, Hyson R, Bertram R. Electrophysiological characterization and computational models of hvc neurons in the zebra finch. Journal of Neurophysiology 110 (2013) 1227 -- 1245. [26] Bolhuis JJ, Okanoya K, Scarff C. Twitter evolution: Converging mechanisms in birdsong and human speech. Nature Reviews Neuroscience 11 (2010) 747 -- 759. [27] Fee MS, Scharff C. The songbird as a model for the generation and learning of complex sequential behaviors. ILAR Journal 51 (2010) 362 -- 377. doi:10.1093/ilar.51.4.362. [28] Mooney R. Neuronal mechanisms for learned birdsong. Learning and Memory 16 (2009) 655 -- 669. doi:10.1101/lm.1065209. [29] Simonyan K, Horwitz B, Jarvis E. Dopamine regulation of human speech and bird song: A critical review. Brain and Language 122 (2012) 142 -- 150. doi:10.1016/j.bandl.2011.12.009. [30] Teramitsu I, Kudo LC, London SE, Geschwind DH, White SA. Parallel foxp1 and foxp2 expression in songbird and human brain predicts functional interaction. Journal of Neuroscience 24 (2004) 3152 -- 3163. doi:10.1523/JNEUROSCI.5589-03.2004. [31] Jarvis ED, Gunturkun O, Bruce L, Csillag A, Karten H, Kuenzel W, et al. Avian brains and a new understanding of vertebrate brain evolution. Nature Review Neuroscience 6 (2005) 151 -- 159. doi:10.1038/nrnl1606. [32] Doupe AJ, Perkel DJ, Reiner A, Stern EA. Birdbrains could teach basal ganglia research a new song. Trends in Neuroscience 28 (2005) 353 -- 363. doi:10.1016/j.tins.2005.05.005. [33] Nottebohm F, Stokes TM, Leonard CM. Central control of song in the canary, serinus canarius. Journal of Comparative Neurology 165 (1976) 457 -- 486. doi:10.1002/cne.901650405. [34] Fortune ES, Margoliash D. Parallel pathways and convergence onto hvc and adjacent neostriatum of adult zebra finches (taeniopygia guttata). Journal of Comparative Neurology 360 (1995) 413 -- 441. doi:10.1002/cne.903600305. [35] Brainard MS, Doupe AJ. Auditory feedback in learning and maintenance of vocal behaviour. Nature Reviews Neuroscience 1 (2000). [36] Hamill OP, Marty A, Neher E, Sakmann B, Sigworth FJ. Improved patch-clamp techniques for high-resolution current recording from cells and cell-free membrane patches. Pflugers Archiv 391 (1981) 85 -- 100. doi:10.1007/BF00656997. [37] Hern´andez-Ochoa EO, Schneider MF. Voltage clamp methods for the study of membrane currents and sr ca2+ release in adult skeletal muscle fibres. Progress in Biophysics and Molecular biology 108 (2012) 98 -- 118. [38] Bagal SK, Marron BE, Owen RM, Storer RI, Swain NA. Voltaged gated sodium channels as drug discovery targets. Channels (Austin) 9 (2015) 360 -- 366. doi:10.1080/19336950.2015.1079674. [39] Frolov RV, Weckstrom M. Ion channels as therapeutic targets, part a. Advances in Protein Chemistry and Structural Biology 103 (2016) 1 -- 386. [40] Daou A, Margoliash D. Things to say about birds; brothers and not. Waiting for Godot 11901 (2018). [41] Breen D. Characterizing Real World Neural Systems Using Variational Methods of Data Assimilation. Ph.D. thesis, University of California San Diego (2017). [42] Wang J, Breen D, Akinin A, Broccard F, Abarbanel HDI, Cauwenberghs G. Assimilation of biophysical neuronal dynamics in neuromorphic vlsi. Biomedical Circuits and Systems 11 (2017) 1258 -- 1270. [43] Kadakia N, Armstrong E, Breen D, Daou A, Margoliash D, Abarbanel H. Nonlinear statistical data assimilation for hvcra neurons in the avian song system. Biological Cybernetics 110 (2016) 417 -- 434. Frontiers 27 Miller et al. [44] Breen D, Shirman S, Armstrong E, Kadakia N, Abarbanel H. Hvci neuron properties from statistical data assimilation (2016). [45] Nogaret A, Meliza CD, Margoiliash D, Abarbanel. Automatic construction of predictive neuron models through large scale assimilation of electrophysiological data. Scientific Reports 6 (2016). doi:10.1038/srep32749. [46] Mooney R, Prather JF. The hvc microcircuit: The synaptic basis for interactions between song motor and vocal plasticity pathways. Journal of Neuroscience 25 (2005) 1952 -- 1964. doi:10.1523/ JNEUROSCI.3726-04.2005. [47] Kosche G, Vallentin D, Long MA. Interplay of inhibition and excitation shapes a premotor neural sequence. J Neurosci. 35 (2015) 1217 -- 1227. [48] Goldman DE. Potential, impedance, and rectification in membranes. The Journal of General Physiology 27 (1943) 37 -- 60. doi:10.1085/jgp.27.1.37. [49] Dutar P, Vu HM, Perkel DJ. Multiple cell types distinguished by physiological, pharmacological, and anatomic properties in nucleus hvc of the adult zebra finch. Journal of Neurophysiology 80 (1998) 1828 -- 1838. doi:10.1152/jn.1998.80.4.1828. PMID: 9772242. [50] Mooney R. Different subthreshold mechanisms underlie song selectivity in identified hvc neurons of the zebra finch. Journal of Neuroscience 20 (2000) 5420 -- 5436. doi:10.1523/JNEUROSCI. 20-14-05420.2000. [51] Kubota M, Taniguchi I. Electrophysiological characteristics of classes of neuron in the hvc of the zebra finch. Journal of Neurophysiology 80 (1998) 914 -- 923. doi:10.1152/jn.1998.80.2.914. PMID: 9705478. This is a provisional file, not the final typeset article 28
1611.00033
1
1611
2016-07-26T06:04:02
Wavelet algorithm for the identification of P300 ERP component
[ "q-bio.NC" ]
Brain-computer interfaces have many algorithms based on the P300 component of ERP. Modern industry has started to produce consumer grade EEG equipment which is handy and not too expensive. This gives us an opportunity to use BCI in everyday practice. In order to improve the performance of these devices, we need effective algorithms for time series analysis and pattern recognition. We have tested Emotiv Insight headset in a real live environment and we have conducted several tests for series of standard wavelets in P300 pattern recognition task.
q-bio.NC
q-bio
Wavelet algorithm for the identification of P300 ERP component © 2016, S.N. Agapov 1, V.А. Bulanov 2, А.V. Zaharov 3, М.S. Sergeeva 4. 1 [email protected] 2 [email protected] IT Universe LLC, Department of BCI. Samara, Russia. 3 [email protected] Samara State Medical University, Department of Neurology and Neurosurgery. Samara, Russia. 4 [email protected] Samara State Medical University, Department of the Normal Physiology. Samara, Russia. Abstract: Brain computer interfaces have many algorithms based on the P300 component of ERP. Modern industry has started to produce consumer grade EEG equipment which is handy and not too expensive. This gives us an opportunity to use BCI in everyday practice. In order to improve the performance of these devices we need effective algorithms for time series analysis and pattern recognition. We have tested Emotiv Insight headset in real live environment and we have conducted several tests for series of standard wavelets in P300 pattern recognition task. Copyright: 2016 IT Universe LLC. This is an open-access article distributed under the terms of the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original authors and source are credited. Competing interests: The authors have declared that no competing interests exist. Wavelet algorithm for the identification of P300 ERP component, © 2016 Agapov et al. p. 1 / 10 1. Introduction Modern population is becoming older every year. In the 20th century the average life expectancy rose significantly [1]. This resulted in bringing about several ageing-related blocking diseases such as amiotrophic lateral sclerosis (ALS), brainstem stroke and others [2, 3]. Heart diseases have become the main cause of death since the middle of the 20th century [4]. They lead to after-effects which can be the cause of complete or partial paralysis. The rehabilitation of these people increases the expenses allocated by a developed country's government on healthcare system [5, 6]. All these factors require fast development of effective and cheap brain-computer interfaces. If BCI equipment is used together with appropriate medical and rehab devices, it can considerably improve the abilities of movement-constrained patients. Nowadays brain-computer interfaces (BCI) can use several different types of input signals from brain: the P300 component of ERP-wave, imagery movements, slow cortical potentials (SCP), steady state visually evoked potential (SSVEP) [7]. Modern industry has produced a lot of devices which use some of these concepts. However, in everyday rehabilitation clinical practice their use is limited. One of the main causes is the cost of the software and hardware parts of the brain-computer interfaces. Low-cost devices often have poor SNR (signal-to-noise ratio) characteristics. The modern methods of digital signal processing and machine learning can greatly improve the accuracy of consumer class equipment. This improvement should give an opportunity to use EEG devices for medical and rehabilitation facilities. The implementation of wavelet analysis for the recognition of P300 ERP component in EEG signal registered with the help of Emotiv Insight (Emotiv Inc., http://emotiv.com/insight/) is shown in this paper. 2. Materials and methods A group of 5 healthy men aged 29-44 (35.8 ± 7.2) participated in our experiments. We recorded EEG signals with 5-channel Emotive Insight EEG headset (Fig. 1). This device uses channels AF3, AF4, T7, T8, Pz according to the international 10-20 system (Fig. 2). The reference electrode was always placed at the left ear. We used software that showed stimulus on the computer display («eSpeller», Java 1.8, Windows 7, granted by CPR LLC, Samara, Russia). That software had been designed with "single character paradigm" [11]. Fig. 1. Emotiv Insight 5-channel wireless EEG headset. Wavelet algorithm for the identification of P300 ERP component, © 2016 Agapov et al. p. 2 / 10 Fig. 2. Electrodes disposition. EEG was recorded in a small office room; all tests were performed when it was daylight. There was some noise made by PC coolers. Datasets were recorded with the sampling rate 128 sec-1. The obtained data was analyzed with MathWork® MATLAB R2015a (http://www.mathworks.com) in Department of BCI of Samara State Medical University, and R statistical package (https://www.r-project.org). 2.1. Experimental design We carried out our experiments within 9 days between 12:00 and 18:00. Each participant sat in front of the PC display. The distance from a participant's eyes to the display was 0.5-0.7 m. The participant wearing Emotiv Insight EEG headset was gazing at the display, when eSpeller software showed him the square window with 250-mm-side size divided by 9 equal cells in the form of 3x3 matrix with a dot in each cell. The software flashed the cells in random order. 'Flashed' means changed the image in the cell: it made the dot one and a half as large and it also altered the color from grey [RGB 179,179,179] to red [RGB 255, 0, 0] (Fig. 3). Each of the 9 cells was flashed only once in each cycle. One cycle consists of 9 flashes. We set stimulus duration to 120 ms and pause duration to 180 ms. Therefore, ISI (interstimulus interval) was set to 300 ms [22]. The 20 flashing cycles completed one experimental session. There was a 10-sec break between the sessions. Fig. 3. The window of eSpeller. Cells 1 – 4, 6 – 9 are not flashing, the red dot is flashing in cell 5. Wavelet algorithm for the identification of P300 ERP component, © 2016 Agapov et al. p. 3 / 10 Each participant had a task to concentrate on the only one cell at the moment. That cell was defined as target cell and the red dot in this cell was called target visual stimulus. Therefore, the other 8 cells were called non-target cells, and the red dots in them were named non-target visual stimulus. The cells were flashed in random order. In the course of the experiment each participant gazed at cells 1-9 in succession (Fig. 4). Fig. 4. The succession in which cells are gazed at. A participant gazes at the cell 1 first, after 20 flashing cycles they have a 10-sec break and then they continue with cell 2. They repeat this procedure up to cell 9. We presented visual stimuli and recorded EEG signals making time stamps simultaneously. As a result, we got 9 files, one for each session, in accordance with the number of cells which were gazed at by a participant. We conducted 3 trials for each participant. All the participants were healthy dexter men and they were volunteers. In this paper the subjA, subjB, subjK, subjP, subjS represent the participants. Therefore, we conducted 15 trials and recorded 135 EEG-signals – 9 files × 3 trials × 5 participants. 2.2. Continuous and discrete wavelet transformations The Fourier transform requires stationarity of the signals and does not locate any events in the signal in time domain. The wavelets give us this opportunity. The wavelets spread in many areas such as time series analysis, image processing, variance stabilization and others. [12, 13]. One of the features of the wavelets is scaling, which variates the width of the time window used for the analysis [14]. A wide class of functions can be represented as wavelets, and then we can create an adapted wavelet from time series data. Up until now many different wavelets have been created and they have a wide area of applications. For example, they are Morlet, Daubechies, "mexican hat" wavelets. The Matlab wavelet toolbox presents 16 distinct wavelet families. The function 𝜓(𝑡) which satisfies the next conditions is called wavelet: 𝜓-𝑢𝑑𝑢=1 𝜓𝑢𝑑𝑢=0 , When these conditions are adhered to, there should be an interval [−𝑇,𝑇] of finite length and the value of 𝜓(𝑡) close to 0 outside of that interval. As a result, we obtain the "small wave" or "wavelet" [15]. We used Daubechies wavelets (db4 and db6) for our work, which are shown in Fig. 5. Wavelet algorithm for the identification of P300 ERP component, © 2016 Agapov et al. p. 4 / 10 ) *) ) *) Fig 5. Daubechies wavelets: db4 (left), db6 (right). The continuous wavelet transform (CWT) of a signal 𝑥𝑡 ∈𝐿-𝑅 is defined as a dot product С7,8 𝜓7,8𝑡 between the signal and the wavelet functions 𝜓7,8(𝑡): С7,8 = <𝑥𝑡,𝜓7,8𝑡 > scaled and translated wavelet function 𝜓(𝑡): 𝑎 𝜓((𝑡−𝑏)/𝑎) 𝜓7,8𝑡 = where 𝑎− scale, 𝑏− translation. the coefficients of the wavelet, The CWT gives coefficients С7,8 which have maximum values at these scales and times where the signal form is similar to that of the wavelet function 𝜓(𝑡). We make CWT with Matlab cwt() function. 𝑎?= 2?, 𝑏?,A= 2?𝑘 . These parameters give us a discrete family of wavelets: 𝜓?,A𝑡 = 2*? 𝜓2*?𝑡−𝑘, 𝑗,𝑘 ∈𝑍 The discrete wavelet transformation (DWT) is defined for discrete scale and translation parameters The discrete wavelet transform adds some more constraints to the function 𝜓(𝑡) [16, 17]. However, we can still perfectly reconstruct a signal 𝑥𝑡 from its discrete wavelet transform coefficients. In the case of the orthogonal wavelets this is: 𝑥𝑡 = 𝐶?,A 𝜓?,A𝑡, 𝑗,𝑘 ∈𝑍 ?,A Wavelet algorithm for the identification of P300 ERP component, © 2016 Agapov et al. p. 5 / 10 Therefore, we can define discrete (dyadic) wavelet transform as: C?,A = <𝑥𝑡,𝜓?,A𝑡 > 3. Results The CWT can be used for the recognition of the patterns which have the form of the wavelet function [18]. When we transform a signal to the wavelet coefficients, the maximum coefficient gives time and scale where the pattern appears. We have analyzed the signals which were recorded during experiment using well-known wavelets such as Daubechies (db 4, 6), reverse biorthogonal wavelets (rbio) and symlets (sym). These wavelets were used in many previous works for EEG filtering and analysis [19 - 21]. One epoch consisted of 128 samples (1 sec), which were counted from the start of visual stimulus. Then we averaged 20 epochs and next we eliminated all linear trends and normalized the averaged signal. For each wavelet we filtered a signal with the same wavelet. For the CWT we used scales 1-64, which covered frequencies 0-30 Hz, and the time interval from 230 ms to 700 ms. It is known from the previous works that P300 component normally appears in that time window [22]. We calculated the total of coefficient's maximum at each time point; this statistic was used for pattern recognition: 𝑆= K 8LM max (С7,8) where С7,8 CWT coefficients, time window width. 𝑤 - We obtained the accuracy of pattern recognition of P300 component for given wavelets. In Tables 1 - 4 we showed the accuracy of the algorithm for different numbers of the averaged epochs. The maximum accuracy for each participant is marked yellow. The maximum accuracy among the participants is marked green. Table 1. P300 detection accuracy (%). The number of the averaged epochs is 5. Participant subjA subjB subjK subjP subjS Db4 14.81 44.44 22.22 25.93 40.74 Db5 11.11 44.44 18.52 40.74 44.44 Wavelet Db6 18.52 44.44 14.81 33.33 44.44 Sym6 14.81 40.74 18.52 25.93 44.44 Table 2. P300 detection accuracy (%). The number of the averaged epochs is 10. Participant subjA subjB subjK subjP subjS Db4 29.63 37.04 37.04 33.33 51.85 Db5 22.22 33.33 37.04 44.44 55.56 Wavelet Db6 18.52 22.22 25.93 40.74 55.56 Sym6 25.93 33.33 37.04 33.33 55.56 Wavelet algorithm for the identification of P300 ERP component, © 2016 Agapov et al. Rbio3.5 14.81 37.04 18.52 33.33 48.15 Rbio3.5 22.22 25.93 37.04 29.63 55.56 p. 6 / 10 Table 3. P300 detection accuracy (%). The number of the averaged epochs is 15. Participant subjA subjB subjK subjP subjS Db4 29.63 44.44 44.44 22.22 59.26 Db5 29.63 51.85 40.74 37.04 70.37 Wavelet Db6 33.33 40.74 37.04 37.04 66.67 Sym6 48.15 44.44 44.44 33.33 59.26 Table 4. P300 detection accuracy (%). The number of the averaged epochs is 20. Particiapnt subjA subjB subjK subjP subjS 4. Discussion Db4 55.56 62.96 51.85 33.33 77.78 Db5 44.44 66.67 51.85 37.04 77.78 Wavelet Db6 48.15 62.96 44.44 37.04 77.78 Sym6 51.85 59.26 51.85 37.04 74.07 Rbio3.5 22.22 40.74 37.04 29.63 70.37 Rbio3.5 40.74 62.96 33.33 37.04 81.48 We can sum up the results obtained from our experiments: - classification accuracy depends strongly on the number of target stimuli. For example, the maximum accuracy for participant subjS changed from 48,15% to 81,48%, when we increased the number of target stimuli from 5 to 20; - classification accuracy strongly depends on a participant. For example, the accuracy for participant subjP was 37,04% and 81,48% for participant subjS (within 20 averaged epochs for both); - classification accuracy depends weakly on the wavelet chosen. For example, the accuracy for participant subjS was 74.07% with wavelet Sym6 and 81,48% with wavelet Rbio3.5; - the average of the classification accuracy was 54,4±7,7% with 20 target stimuli. The minimum was 33,33% for participant subjP with wavelet Db4; the maximum was 81,48% for participant subjS with wavelet Rbio3.5. We can conclude that the current realization of wavelet algorithm for patter recognition in EEG signals has several drawbacks: a low level of generalization and low overall accuracy even with a large quantity of target stimuli. We think there are several reasons for this: - there are anthropomorphic differences between subjects. The shape and time localization of the ERP-waves and their P300 component have high variance among humans. Our results confirm that well- known fact; - the wavelets chosen, in general, have a shape similar to ERP-wave, but this is not a sufficient condition for detection. We can suppose that the adapted wavelet will show a better performance; - we can presume that the Emotiv Insight device has hardware or software design features which make an effect on the given algorithm and do not allow us to get a better result. Wavelet algorithm for the identification of P300 ERP component, © 2016 Agapov et al. p. 7 / 10 5. Conclusion We have presented an algorithm for the P300 component detection in an EEG signal using several well-known wavelets. This algorithm can be used with Emotiv Insight device for a certain category of people. Additional testing of the accuracy level for a certain participant is required before using the given algorithm. Accuracy is likely to be improved if the adapted wavelet is used, and we are planning this investigation in the future. The results obtained need more investigation. For example, we can test the given algorithm on the data acquired with other EEG equipment. Also, we could try to adapt a wavelet to the individual pattern of the ERP wave of a participant. Wavelet algorithm for the identification of P300 ERP component, © 2016 Agapov et al. p. 8 / 10 References [1] [2] [3] [4] [5] [6] [7] [8] [9] J. Oeppen, J. W. Vaupel, and others, "Broken limits to life expectancy," Science (80-. )., vol. 296, no. 5570, pp. 1029–1031, 2002. G. Logroscino, R. Tortelli, G. Rizzo, B. Marin, P. M. Preux, and A. Malaspina, "Amyotrophic lateral sclerosis: An aging-related disease," Curr. Geriatr. Reports, vol. 4, no. 2, pp. 142–153, 2015. B. a Yankner, T. Lu, and P. Loerch, "The aging brain.," Annu. Rev. Pathol., vol. 3, pp. 41–66, 2008. [Online]. Available: http://www.who.int/mediacentre/factsheets/fs310/ru/. K. McGregor and B. Pentland, "Head injury rehabilitation in the U.K.: An economic perspective," Soc. Sci. Med., vol. 45, no. 2, pp. 295–303, 1997. C. D.N. and O. J., "A clinical and economic perspective on head injury rehabilitation," Journal of Head Trauma Rehabilitation, vol. 8, no. 4. pp. 1–14, 1993. G. Edlinger, B. Z. Allison, and C. Guger, "How many people can use a BCI system?," in Clinical Systems Neuroscience, 2015, pp. 33–66. R. Quian Quiroga, "Obtaining single stimulus evoked potentials with wavelet denoising," Phys. D Nonlinear Phenom., vol. 145, no. 3–4, pp. 278–292, 2000. R. Quian Quiroga, O. W. Sakowitz, E. Basar, and M. Schürmann, "Wavelet Transform in the analysis of the frequency composition of evoked potentials," Brain Res. Protoc., vol. 8, no. 1, pp. 16–24, 2001. [10] X. Li, X. Chen, Y. Yan, W. Wei, and Z. Wang, "Classification of EEG Signals Using a Multiple Kernel Learning Support Vector Machine," Sensors, vol. 14, no. 7, pp. 12784–12802, 2014. [11] R. Fazel-rezai, S. Gavett, W. Ahmad, A. Rabbi, and E. Schneider, "A Comparison among Several P300 Brain-Computer Interface Speller Paradigms," Clin. EEG Neurosci., vol. 42, no. 4, pp. 209– 213, 2011. [12] C. Heil, "Ten Lectures on Wavelets (Ingrid Daubechies)," SIAM Review, vol. 35, no. 4. pp. 666– 669, 1993. [13] S. Mallat, "A Wavelet Tour of Signal Processing," A Wavelet Tour Signal Process., pp. 20–41, 1999. [14] C. K. Chui, Wavelets : a tutorial in theory and applications, no. 2. 1992. [15] D. B. Percival and A. T. Walden, "Wavelet Methods for Time Series Analysis," J. Nonparametr. Stat., vol. 10, no. December 2014, p. 594, 2000. [16] E. Erçelebi, "Electrocardiogram signals de-noising using lifting-based discrete wavelet transform," Comput. Biol. Med., vol. 34, no. 6, pp. 479–493, 2004. [17] M. V. Wickerhauser, A. Jensen, A. la Cour-Harbo, A. Boggess, and F. J. Narcowich, Ripples in Mathematics: The Discrete Wavelet Transform, vol. 110, no. 2. 2003. [18] D. B. Percival and A. T. Walden, Wavelet Methods for Time Series Analysis. 2000. [19] A. B. Geva and D. H. Kerem, "Forecasting generalized epileptic seizures from the EEG signal by wavelet analysis and dynamic unsupervised fuzzy clustering," IEEE Trans. Biomed. Eng., vol. 45, no. 10, pp. 1205–1216, 1998. Wavelet algorithm for the identification of P300 ERP component, © 2016 Agapov et al. p. 9 / 10 [20] V. J. Samar, A. Bopardikar, R. Rao, and K. Swartz, "Wavelet Analysis of Neuroelectric Waveforms: A Conceptual Tutorial," Brain Lang., vol. 66, no. 1, pp. 7–60, 1999. [21] M. Murugappan, "Classification of human emotion from EEG using discrete wavelet transform," J. Biomed. Sci. Eng., vol. 03, no. 04, pp. 390–396, 2010. [22] S. J. Luck, An Introduction to the Event-Related Potential Technique, vol. 78, no. 3. 2005. Wavelet algorithm for the identification of P300 ERP component, © 2016 Agapov et al. p. 10 / 10
1709.03978
1
1709
2017-09-11T22:39:35
A neuronal network model of interictal and recurrent ictal activity
[ "q-bio.NC" ]
We propose a neuronal network model which undergoes a saddle-node bifurcation on an invariant circle as the mechanism of the transition from the interictal to the ictal (seizure) state. In the vicinity of this transition, the model captures important dynamical features of both interictal and ictal states. We study the nature of interictal spikes and early warnings of the transition predicted by this model. We further demonstrate that recurrent seizures emerge due to the interaction between two networks.
q-bio.NC
q-bio
A neuronal network model of interictal and recurrent ictal activity M. A. Lopes,1, 2, 3, 4, ∗ K.-E. Lee,4, 5 and A. V. Goltsev4, 6 1College of Engineering, Mathematics and Physical Sciences, University of Exeter, Devon EX4, United Kingdom 2Wellcome Trust Centre for Biomedical Modelling and Analysis, University of Exeter, Devon EX4, United Kingdom 3EPSRC Centre for Predictive Modelling in Healthcare, University of Exeter, Devon EX4, United Kingdom 4Department of Physics & I3N, University of Aveiro, 3810-193 Aveiro, Portugal 5Department of Anesthesiology and Center for Consciousness Science, University of Michigan Medical School, Ann Arbor, Michigan, USA 6A.F. Ioffe Physico-Technical Institue, 194021 St. Petersburg, Russia We propose a neuronal network model which undergoes a saddle-node on an invariant circle bifurcation as the mechanism of the transition from the interictal to the ictal (seizure) state. In the vicinity of this transition, the model captures important dynamical features of both interictal and ictal states. We study the nature of interictal spikes and early warnings of the transition predicted by this model. We further demonstrate that recurrent seizures emerge due to the interaction between two networks. PACS numbers: 87.19.lj,87.19.xm,05.70.Fh,05.10.-a,87.19.ln,87.18.Sn I. INTRODUCTION Epilepsy affects nearly 1% of the population world- wide [1]. This disabling neurological disorder is char- acterized by spontaneous recurrent seizures, which cor- relate to strongly synchronized neuronal activities, so- called paroxysmal activity, revealed in electroencephalo- grams (EEG). Anticonvulsant medications can prevent seizures, but side effects are frequently reported [2]. For about 30% of the patients, medications are not effec- tive [3]. A minority undergoes surgery to remove the epileptogenic brain tissue, but even in these cases the patients may continue experiencing spontaneous seizures [4, 5]. One of the main challenges has been to try to fore- cast seizures [6 -- 8]. On one hand, the unpredictability of seizure occurrence is a major burden of the condition [9] and therefore being able to alert patients of impending seizures could greatly improve their quality of life. On the other hand, it would allow the design of closed-loop inter- vention systems which could stop seizures [10]. Much of the research in seizure prediction has been focused on al- gorithms [8], however a better understanding of epilepsy mechanisms is required. Although epilepsy is an umbrella term for a range of syndromes, the electrophysiological signatures are similar between them [11]. For instance, different epileptogenic lesions can produce similar electroencephalographic pat- terns [12]. Also, it is noteworthy that it is possible to in- duce seizures in non-epileptic brains across species both in vivo and in vitro, which again present similar electro- physiological features (see e.g. [13, 14]). As argued by Jirsa et al. [11], these facts suggest the existence of invari- ant dynamical properties underlying seizure dynamics. ∗ [email protected] Moreover, there are evidences that seizures self-terminate via a critical transition [15]. Note that bifurcations are the mechanisms of phase transitions in many-body in- teracting systems [16, 17]. The two frameworks provide complementary insights about the underlying transitions. While there are extensive studies on what kind of bi- furcations occur at the onset and offset of seizures (e.g. [11, 18, 19]), there are fewer studies about the collective nature of the phase transitions [20]. The fundamental question is how does the spike interaction between large populations of neurons may result in seizures. Critical phenomena provide early-warnings of phase transitions [17], which consequently open the possibility to take action to prevent the occurrence of those transi- tions. We have previously demonstrated that the inter- action between neurons on a network give rise to collec- tive phenomena and diverse phase transitions [21]. Note that different phase transitions are associated with dif- ferent precursors and different critical phenomena such as (among other) bursts of neuronal activity, avalanches, hysteresis, critical slowing-down, symmetry breaking, and resonance phenomena [21, 22]. Here, we propose a neuronal network model consist- ing of interacting excitatory and inhibitory neurons to understand the nature and emergence of both interictal and ictal activity. We start by presenting the model and its dynamical states. We then study the properties of interictal-like spikes, which are evoked in the vicinity of the transition to the ictal state. As the dynamical state moves towards the transition, early-warning phenomena signal the impending transition. We demonstrate that such phenomena is revealed through the stimulation of interictal-like spikes, and the analysis of accompanying low-fluctuating activity. Finally, we show that recurrent ictal activity is an emergent collective phenomenon in a system of two interacting neuronal networks. II. NEURONAL NETWORK MODEL Herein we consider a neuronal network model [21 -- 23], which we will refer to as the stochastic cellular automata neuronal network model (SCANNM). In the SCANNM, neurons are modelled as stochastic integrate-and-fire neu- rons: they integrate the inputs and fire a train of spikes with a certain probability if the input is larger than an activation threshold (see Appendix A for more details). There are two populations of neurons, excitatory and in- hibitory neurons. Excitatory neurons fire positive out- puts, whereas inhibitory fire negative outputs to their postsynaptic neighbours. In general, the two populations have different response times to stimuli. Here we consider the case in which excitatory neurons respond faster than inhibitory neurons. Additionally, neurons are excited by endogenous stimuli that account for random spikes com- ing from other areas of the brain, as well as spontaneous releases of neurotransmitters at the synapses. (cid:104)n(cid:105) is the endogenous stimulation and we use it as control parame- ter. The neurons form an uncorrelated random directed network (properties of this kind of networks have been studied, for example, in [24 -- 26]). The mean-field neu- ronal dynamics are described by the fractions of active excitatory (ρe) and inhibitory (ρi) neurons which follow the rate equations [21, 23] ρa µa = −ρa + Ψa(ρe, ρi,(cid:104)n(cid:105)), (1) where a = e, i, and ρ ≡ dρ/dt. Ψa(ρe, ρi,(cid:104)n(cid:105)) is the probability that at time t a randomly chosen excitatory (a = e) or inhibitory (a = i) neuron becomes active. This function encodes the network structure, single neu- ron stochastic firing rules, and endogenous stimulation [21 -- 23]. The model is analytically solvable, but despite its sim- plicity, it describes a rich repertoire of collective phenom- ena, namely neuronal avalanches, bursty activity, hys- teresis, bistability, different kinds of neuronal oscillations, phase transitions, and stochastic resonance [21, 22]. Fur- thermore, the SCANNM combines two usually distinct modelling frameworks to describe mesoscopic brain dy- namics, namely it allows the modelling of large-scale neu- ronal networks like in Ref. [27], and it is simultaneously described by a neural mass formulation (see for instance [28]). It thus enables an analysis of both single neuron dynamics within the network, and large-scale dynamics of neuronal populations that can be treated numerically and analytically. Figure 1 depicts the different patterns of neuronal ac- tivity in the SCANNM. In this paper we focus on the re- gions corresponding to (i) low fluctuating activity around a stable state, which we identify as a 'normal' state; (ii) low fluctuating activity with sporadic single sharp os- cillations, the 'interictal' state; and (iii) sustained net- work oscillations, which is the model 'ictal' state. The boundary between the interictal and ictal regions corre- sponds to a saddle-node on an invariant circle (SNIC) 2 FIG. 1. Schematic diagram of the different dynamical states of the SCANNM as function of the endogenous stimulation. We find low neuronal activity at low stimulation ((cid:104)n(cid:105) < nc1). At an intermediate stimulation level (nc1 < (cid:104)n(cid:105) < nc2), the neuronal network exhibits transients of high activity, either bursts or single sharp oscillations, which give place to damped oscillations or sustained oscillations at (cid:104)n(cid:105) > nc2 respectively depending on parameters. The shape of the sustained oscilla- tions changes from slow, high-amplitude oscillations to fast, low-amplitude oscillations as the stimulation increases. The neuronal network produces damped oscillations to a high ac- tivity state at (cid:104)n(cid:105) > nc3. bifurcation (at (cid:104)n(cid:105) = nc2), which is the critical point of a second-order phase transition from the interictal to the ictal state [21]. At higher endogenous stimulations, the ictal region is bounded by the critical point of a supercrit- ical Hopf bifurcation that separates the ictal state from a high activity state (at (cid:104)n(cid:105) = nc3). In the vicinity of the Hopf bifurcation, the neuronal oscillations have high fre- quency, and low amplitude, in contrast to the oscillations close to the SNIC bifurcation, which are characterized by low-frequency and high-amplitude [29]. In the interictal state, interictal-like spikes (ILS) emerge at random, but with a deterministic shape. Note that ILS are strongly nonlinear events that comprise the synchronous activity of almost 90% of the neurons in the network. An ILS is described by a trajectory that goes around an unsta- ble point in the (ρe, ρi)−phase plane (see Sec. III). Their occurrence is deterministic if the activity overcomes a threshold (a separatrix in the phase plane). This thresh- old defines the number of excitatory neurons that must be activated simultaneously in order to generate an ILS. As an example, for a network of 104 neurons at (cid:104)n(cid:105) = 16 (i.e., at (nc2−(cid:104)n(cid:105))/nc2 ≈ 0.15), which is in a low activity state where almost all neurons are inactive, the simultaneous activation of just 75 excitatory neurons chosen at random (i.e., about 1% of the excitatory neurons in a network with 25% inhibitory neurons) generates an ILS formed by the synchronized activity of about 9000 neurons [21]. One should note that the duration of these ILS is much larger than the period of single neuron spikes. This im- plies that an ILS is genuinely a collective phenomena in low activitity(normal state)single sharposcillations(interictal spikes)bursting activity(bistable state)dampedoscillations(spindles)slow sharposcillations(ictal activity)fast oscillationsnc1c2c3nn spite of the fact that it can be elicited by a small num- ber of neurons. For realistic parameters, namely mean degree c = 1000, fraction of inhibitory neurons gi = 0.25, synaptic efficacies ratio Ji/Je = −3, dimensionless fir- ing threshold Ω = 30, ratio between excitatory and in- hibitory response times α = µi/µe = 0.7, and 1/µe = 20 ms (see Appendix A and Refs. [21 -- 23]), the typical du- ration of an ILS is about 100 ms which is comparable to real interictal spikes [30]. Within the ictal region, the frequency of sustained oscillations increases with endoge- nous stimulation from very low frequencies up to 4 Hz [21], which is comparable to the frequency of ictal activ- ity [31]. These are natural features of the model, without needing to calibrate parameters. III. THE NATURE OF THE INTERICTAL-LIKE SPIKES In order to understand the nature of ILS, let us study their phase trajectories in the plane ρe − ρi. ILS emerge in the region nc1 < (cid:104)n(cid:105) < nc2 (see Fig. 1). In this re- gion there are three fixed points: a stable fixed point at low activity, a saddle point at an intermediate activ- ity, and an unstable point at high activity [21]. For this parameter region, the dynamics in the SCANNM is qual- itatively equivalent to the dynamics of the Morris-Lecar neuron near the SNIC bifurcation [32]. To the best of our knowledge, a SNIC bifurcation has not been found in any other neuronal network model. Thus, the SCANNM gives an unique possibility to study the collective behav- ior of neuronal populations near a SNIC bifurcation. We follow the nomenclature used by Rinzel and Ermentrout [32]: the stable point corresponds to a rest state (R), and the saddle point is a threshold (T) on the separatrix that divides the phase plane into two regions. There are two heteroclinic orbits connecting T to R, one corresponding to immediate exponential relaxation to R, and another that goes around the unstable point (U), reaching high activity before relaxing to R. The ILS follows the second path (see Fig. 2) as spikes of single neurons do in the Morris-Lecar model. Panels (a) and (b) of Fig. 2 also portray the nullclines of the system, ρa = 0 ⇔ ρa = Ψa(ρe, ρi,(cid:104)n(cid:105)). (2) The nullclines determine the maximums or minimums of excitatory and inhibitory activity. In this case, the ac- tivity of the population a increases ( ρa > 0) below the respective nullcline, whereas it decreases above. Con- sequently, the ILS move counterclockwise in the phase plane. Thus, any excitatory activity perturbation that drives the activity state below both nullclines results in an ILS as the one displayed in Fig. 2(c). At the critical point nc2, the points R and T merge, and there is a ho- moclinic orbit around the unstable point connecting the saddle-node to itself. The distance between R and T defines an activation threshold Ath((cid:104)n(cid:105)) for the generation of ILS. We demon- 3 (a) Schematic representation of the nullclines and FIG. 2. fixed points in the (ρe, ρi)−phase plane. (b) Nullclines and fixed points of the SCANNM (numerical integration of Eqs. (2)). In panels (a) and (b) the blue and red lines corre- spond to the ρe- and ρi-nullcline, respectively. The nullclines intersect at the fixed points: the stable or rest state (R), the saddle or threshold (T), and the unstable point (U). (c) Tra- jectory of an ILS in the (ρe, ρi)−phase plane. The line is the result of the numerical integration of Eqs. (1) and the dots represent simulations of the model. Parameters: c = 1000, Ω = 30, gi = 0.25, Ji = −3Je, σ2 = 10, (cid:104)n(cid:105) = 16, α = 0.7, τ = 0.1 and N = 104 (see the Appendix A for details about the parameters). strate in Appendix B that Ath((cid:104)n(cid:105)) follows a square root dependence with (cid:104)n(cid:105) in the vicinity of the SNIC bifurca- tion, Ath((cid:104)n(cid:105)) ∝(cid:112)nc2 − (cid:104)n(cid:105). (3) In a finite network, finite-size effects elicit activity fluctu- ations which can overcome the activation threshold pro- vided that the system is sufficiently close to the critical point, which in turn results in the occasional generation of ILS (see Fig. 3). IV. EARLY WARNINGS OF THE TRANSITION TO THE ICTAL STATE If a control parameter such as the endogenous stimu- lation (cid:104)n(cid:105) changes slowly from the normal state towards 4 FIG. 3. (a) Series of ILS generated at random due to finite- size effects at (cid:104)n(cid:105) = 18.75 (nc2 = 18.8). (b) Zoom of an ILS from panel (a). All ILS have the same shape. Parameters in simulations are the same as in Fig. 2. FIG. 4. Panels (a) and (b) depict the minimum signal am- plitude Fmin to elicit an ILS and the minimum time between two consecutive ILS Dmin as function of the endogenous stim- ulation (cid:104)n(cid:105), respectively. Parameters in the numerical inte- gration of the rate equations (1) are the same as in Fig. 2. the ictal state through the interictal state, changes in the neuronal dynamics can inform on how close is the system to the critical point of the transition to the ictal state. However, if the variation of the control param- eter is too fast to observe its consequences in the dy- namics, the transition cannot be anticipated [33]. In the SCANNM, since the threshold decreases towards the crit- ical point (Eq. (3)), finite-size effects are more likely to generate ILS, and therefore the rate of ILS is expected to increase. However, as the neuronal network approaches the critical point nc2, the stable point R gets closer to the saddle point T. The time D that the system spends in this region of the phase plane increases as D ∝ (nc2 − (cid:104)n(cid:105))−1/2. (4) This equation describes a general feature of a SNIC bi- furcation [16]. Therefore, although it becomes easier to elicit ILS from the rest state as the system approaches the critical point, the refractory time also increases, i.e., it takes longer to invoke consecutive ILS. To study the generation of ILS as function of the distance to nc2, we stimulate the neuronal network with an excitatory force. In Eqs. (1) we introduce an additional term (1 − ρe)F that stimulates the inactive excitatory population (1−ρe) with a delta-like field of amplitude F (duration equal to a time step). We found the minimum amplitude Fmin required to elicit ILS as function of endogenous stimula- tion, and measured the minimum time between two con- secutive ILS Dmin generated by two consecutive pulses Fmin. Figure 4 shows that Dmin diverges as the system approaches nc2, as Eq. (4) predicts, and Fmin tends to zero. This is because Fmin is essentially a measure of the activation threshold, and as such Fmin ∝(cid:112)nc2 − (cid:104)n(cid:105). (5) For a sufficiently small fixed stimulation F , the neuronal network generates ILS if F ≥ Fmin((cid:104)n(cid:105)). The stimulation dependence of Dmin is a consequence of critical slowing down near nc2. (Note that "critical slowing down" is a phenomenon by which a system takes longer and longer time to recover from small perturbations as it approaches a critical point of a continuous phase transition [17].) At the critical point, Fmin → 0 and D → ∞, meaning that network oscillations emerge with zero frequency, which corresponds to the homoclinic orbit mentioned above. FIG. 5. Zero-frequency power enhancement as a precursor of the transition to the ictal state. Panels (a) and (b) show the power spectral density of activity fluctuations in decibel (10 log10(PSD)) at (cid:104)n(cid:105) = 18.6 and (cid:104)n(cid:105) = 18.8, respectively. (c) Maximum of the power spectral density at zero frequency as function of the endogenous stimulation. Parameters in simulations are the same as in Fig. 2. Besides these dynamical changes involving ILS, the low activity state is also affected by the critical slowing down. This can be quantified by power spectral analysis of low activity fluctuations near the critical point. Using the Wiener-Khintchine theorem, we have demonstrated that the power spectral density (PSD) of activity fluctuations in the low activity state when (cid:104)n(cid:105) → nc2 has a sharp zero-frequency peak which grows as Smax ∝ 1/(nc2−(cid:104)n(cid:105)) [21]. This behavior was demonstrated in a metastability region in the vicinity of a first-order phase transition in Ref. [21], and it also occurs near the second-order phase transition under consideration. Figure 5 shows that the zero-frequency peak of the PSD increases as the neuronal network approaches the critical point nc2. 0 100000t 0 1 (a)89458970t01(b)1617181950 350(b)16171819 0 0.4(a) 0 100200f /Hz-150-100 -50PSD /dB(a) 0 100200f /Hz-150-100 -50PSD /dB(b) 18 18.518.902468max(PSD)10-3(c) 5 FIG. 6. Ictal-like pattern evolution driven by endogenous stimulation in the SCANNM. The blue line is time-dependent stimulation (cid:104)n(cid:105) that varies in the range [16, 50] (right y-axis) and drives the neuronal activity ρ (black line) from the low ac- tivity state to fast low-amplitude oscillations with a DC shift, which then evolves to high-amplitude, low-frequency oscilla- tions, before returning to the low activity state. Parameters in the numerical integration of the rate equations (1) are the same as in Fig. 2. V. MODEL OF RECURRENT TRANSITIONS TO THE ICTAL STATE If we relax the condition of a slowly changing control parameter, the SCANNM is capable of mimicking a typ- ical ictal pattern evolution [11, 31]. Figure 6 shows that an abrupt increase of the endogenous stimulation can bring the neuronal network from the low activity state to the vicinity of the supercritical Hopf bifurcation, which results in a DC shift of the neuronal activity accompa- nied by fast low-activity oscillations. As then (cid:104)n(cid:105) slowly decreases back to the 'normal' or the interictal state, the frequency of the sustained network oscillations decreases, whereas its amplitude increases. At fixed parameters, the SCANNM can either be in the normal, interictal, or ictal state. Recurrent transitions between these states can either be achieved by a change in parameters, or due to external stimuli. Another sce- nario is to consider a network of networks, i.e., several interacting neuronal networks. In general, a network of networks can consist of multiple networks whose inter- network connections can be both directed or undirected, connecting different numbers of excitatory and inhibitory neurons. We consider two interacting neuronal networks as our minimal model of different interacting brain ar- eas. If one of the networks is in the interictal state, then a small additional excitatory input from another network can induce a transition to the ictal state. Recurrent tran- sitions will occur as a consequence of a recurrent input. Such intermittent input can be generated by a network in the bursting state (see Fig. 1). As previously described in Ref. [21], the SCANNM produces recurrent irregular bursts of neuronal activity when close to a first-order phase transition. To illustrate the concept, we consider two networks A and B of size N = 104. Network A is in the bursting state [((cid:104)n(cid:105), α) = (18.7, 0.85), see the Appendix A for the meaning of α] and network B is in the interictal state [((cid:104)n(cid:105), α) = (18, 0.7)]. For simplicity, network A sends axonal projections to network B, but FIG. 7. Minimal model of recurrent ictal transitions. (a) Schematic representation of two networks, where the red net- work (A) influences the blue (B). (b) Neuronal activity in net- work A. (c) Neuronal activity in network B. The clusters of ictal-like activity in network B are driven by the noisy bursts of network A. Panels (d) and (e) are zooms of the activity displayed in panels (b) and (c) respectively. Parameters in simulations are the same as in Fig. 2, except for those re- ferred in the text. network B does not influence network A (see Fig. 7(a)). We define a fraction gAB = 0.3 of excitatory neurons chosen at random in each network, and we wire them by directed connections from A to B (synaptic efficacies JAB = 3Je, see the Appendix A). In Figs. 7(b) and (c), we show the neuronal activity of the two networks. The recurrent transitions to the ictal state in network B are driven by the bursting activity in network A. VI. DISCUSSION In this paper we proposed a neuronal network model to describe interictal spikes and recurrent ictal activity. ILS are strongly nonlinear collective events that comprise the synchronized activity of a large number of neurons. ILS emerge from a low background activity when either random fluctuations or stimuli force the neuronal activ- ity to overcome a threshold (see Fig. 2). This thresh- old becomes smaller as a control parameter (an endoge- nous stimulation) moves the neuronal network towards 0 1000t010 1650 the transition to the ictal state. The transition corre- sponds to a SNIC bifurcation at which sustained net- work oscillations emerge with low frequency. This re- gion of oscillations is also bounded at higher endogenous stimulations by a supercritical Hopf bifurcation. Near the Hopf bifurcation, oscillations have low-amplitude and high-frequency. It is conceivable that the transition to seizures may ei- ther be a consequence of a gradual or an abrupt change in the endogenous stimulation depending on the type of epilepsy [33]. In the SCANNM, if we assume a grad- ual increase of the endogenous stimulation, then we can observe critical phenomena that signal the transition. This transition can be compared to the high amplitude slow (HAS) activity onset pattern observed in some focal epilepsies (see [34] and references therein). We showed that under this assumption, the ILS activation thresh- old scales as the square root of the distance to the SNIC bifurcation, whereas the minimum time between two con- secutive spikes diverges at the critical point of the tran- sition to the ictal state (see Fig. 4). Additionally, the zero-frequency peak of the PSD of low activity fluctua- tions reaches a maximum at the critical point (see Fig. 5). Alternatively, if we instead assume that the control parameter may change abruptly, then the SCANNM is capable of mimicking the typical pattern evolution ob- served in seizures (called low amplitude fast (LAF) ac- tivity onset pattern in [34]). Seizures are often preceded by a low-voltage, high-frequency discharge [35], and the ictal pattern generally displays increasing amplitude and decreasing frequency [31]. The SCANNM exhibits such pattern evolution if we assume that at seizure onset the endogenous stimulation abruptly increases, forcing the neuronal network to jump from the interictal-like state to the vicinity of the Hopf bifurcation. Then, as the stim- ulation gradually decreases towards 'normal' levels, the neuronal activity evolves from low-amplitude, fast oscil- lations, to high-amplitude, slow oscillations (see Fig. 6). Finally, we demonstrated the viability of modelling re- current ictal transitions using two interacting neuronal networks, where the intermittent output of one network drives the other to recurrent seizures. This concept aligns well with other modelling approaches which have also explored the role of interacting populations to generate ictal-like activity [36 -- 39]. Although for simplicity we have considered here the mechanism to seizures as an excitatory drive from another network, we would like to note that this is equivalent to an interruption of inhibi- tion. Such mechanism of triggering seizure-like activity has been observed in genetically engineered mice, where the shut-down of CA2 output leads to hyperexcitability in the recurrent CA3 network [40]. Note however that the networks involved in the generation of ictal activity may be located in distant regions of the brain. This high- lights the importance of studying large scale brain net- works, rather than focal brain activity even in the case of focal epilepsies [41]. Contrarily to previous computational models of 6 epilepsy [11, 42, 43], the SCANNM was not explicitly designed to describe epileptiform activity [21, 23]. In- stead, this model demonstrates that interictal and ictal- like activity may be emergent phenomena of an inter- acting neuronal network. The heterogeneous mean-field equations of the SCANNM were derived from a minimal set of fundamentals, namely neurons behave as stochastic integrate-and-fire neurons, there are two types of neurons (excitatory and inhibitory), and the neurons interact on a complex network [21, 23, 44]. Our analytical and nu- merical results were in good agreement with simulations of the model for sufficiently large networks (typically for N > 103 [23]). We conclude that the SCANNM thus al- lows to study numerically and analytically neural mass- like equations, and to measure and compare single neu- ronal activity. It enables to examine how the activity of single neurons can impact on the whole network. For the case in point, ILS are a remarkable example of a collective network phenomenon that can be evoked by the simul- taneous activation of a few neurons. On the other hand, once an ILS is excited, the SCANNM predicts that it is very difficult to suppress it. The only way to make the neuronal network return to the low activity state is by applying a strong inhibitory stimulus to a considerable macroscopic part of the network. A recent study has reported apparently self- contradictory evidence on the role of pre-ictal spikes for the prediction of seizures [30], showing that different seizures could be preceded by an increase or decrease of the pre-ictal spike rate. The SCANNM provides a possi- ble explanation for this observation. In the model, as a transition to the ictal state is approached, two competing mechanisms can influence the spike rate. On one hand, the activation threshold of ILS decreases which leads to a higher spike rate. On the other hand, critical slowing down hinders the consecutive emergence of spikes (see Fig. 4). It is then conceivable that the prevailing mech- anism may vary from seizure to seizure, and as a result the spike rate can increase or decrease before a seizure. There is also conflicting evidence as to which it may or may not be possible to predict seizures based on critical phenomena [45] or using other data analysis [8]. Nev- ertheless, based on the SCANNM we can propose two measures to forecast seizures. First, Fig. 4(a) indicates that the required external stimulation to evoke ILS be- comes smaller as the neuronal network approaches the transition to the ictal state. In the case of photosensitive epilepsy, the stimulation can be intermittent photic stim- ulation. In other epilepsies it may be necessary to use im- planted electrodes to electrically stimulate the brain, like it was proposed by Silva et al. [46]. Thus, for a given patient, and after a sufficient number of trials, it may be possible to correlate the minimum required stimula- tion to elicit ILS with the timing of impending seizures. However, such method may be infeasible due to the risk of inducing seizures due to this probing stimulation [47]. The second measure does not require stimulation, instead it uses the analysis of ongoing EEG recordings. Figure 5 shows that the power of low frequencies should increase in EEG recordings when approaching a transition to a seizure. In fact, a gradual increase in power of low fre- quencies has been observed as a precursor of spike-wave discharges in absence epilepsy both in humans and rat models [48 -- 51]. We acknowledge, however, that even if this process takes place, it may often be unobservable be- cause if the transition occurs in a faster time scale than the scale of the low frequencies, then it is not possible to find the gradual power increase. VII. CONCLUSION In this paper we proposed a neuronal network model (the SCANNM) to describe interictal and ictal activ- ity, as well as ictal-like pattern evolution, and spon- taneous recurrent transitions to seizures. Additionally, we found a set of precursors that signal the transition to the ictal state. The neuronal activity state was de- pendent of an endogenous stimulation which we used as the control parameter. The interictal state was char- acterized in the model by low fluctuating activity from which interictal-like spikes could sporadically emerge. We demonstrated that the required stimulation to elicit interictal-like spikes tends to zero as the neuronal net- work approaches the critical point of a saddle-node bi- furcation. Furthermore, the transition is signaled by an increase of the zero-frequency peak of the power spec- trum of low activity fluctuations when the endogenous stimulation varies slowly. On the other hand, for abrupt changes in the control parameter, we showed that the model can mimic a typical ictal pattern evolution: as onset with low-voltage, high-frequency discharges, fol- lowed by increasing amplitude, decreasing frequency os- cillations. Finally, we showed that the model could also reproduce recurrent transitions to the ictal state at fixed parameters, as the result of the interaction between two neuronal networks. VIII. ACKNOWLEDGEMENTS This work was partially supported by FET IP Project MULTIPLEX 317532. A.V.G. is grateful to LA I3N for Grant No. PEST UID/CTM/50025/2013. M.A.L. ac- knowledges the financial support of the Medical Research Council (MRC) via grant MR/K013998/01. K.E.L. was supported by the Department of Anesthesiology at the University of Michigan (Ann Arbor) and National In- stitutes of Health (Bethesda, MD, USA) grant RO1 GM098578. Appendix A: SCANNM Here we describe the SCANNM [21 -- 23]. 7 1. Network structure and stochastic dynamics The neuronal network is composed of N stochastic neu- rons, geN excitatory and giN inhibitory (ge + gi = 1). We consider that the network has the structure of the Erdos-R´enyi network. This is a random network with small world properties, namely small mean shortest path length like real neuronal networks in the brain [52]. The neurons are connected by directed edges which repre- sent synapses that allow active neurons to send spikes to their postsynaptic neighbors. In addition, neurons also receive random spikes from endogenous stimulation that represent spontaneous releases of neurotransmitters in synapses and random spikes arriving from other areas of the brain (this stimulation has properties of shot noise [21]). The dynamics of the stochastic neurons is determined by the following rules. If during an integration time win- dow τ the total input Vj(t) to an inactive neuron be- comes larger than a threshold Ω, then with probability τ µa the neuron becomes active and fires a spike train at a constant frequency ν (the index a = e if the neuron is excitatory and a = i if it is inhibitory). If the total input Vj(t) to an active excitatory (inhibitory) neuron becomes smaller than Ω, then the neuron stops to fire with proba- bility τ µa. In this model, the rates µe and µi are the re- ciprocal first-spike latencies of excitatory and inhibitory neurons, respectively. We define a parameter α as the ra- tio of the first-spike latency of excitatory neurons to the first-spike latency of inhibitory neurons, α ≡ µi/µe. If α < 1, then excitatory neurons respond faster to stimuli than inhibitory neurons. This neuronal stochastic behav- ior is meant to account for intrinsic noise within neurons [53], namely, ion channel stochasticity [54]. 2. Rate equations The fractions ρe(t) and ρi(t) of active excitatory and inhibitory neurons, respectively, characterize the neu- ronal activity at time t. They are determined by the rate equations (1), in which Ψa(ρe, ρi,(cid:104)n(cid:105)) is the prob- ability that, at time t, the total input to a randomly chosen excitatory (a = e) or inhibitory (a = i) neu- ron is at least the threshold Ω at a given endogenous stimulation (cid:104)n(cid:105). The functions Ψa(ρe, ρi,(cid:104)n(cid:105)) are deter- mined by the network structure, the distribution func- tion of endogenous stimulation (we consider the Gaus- sian distribution), and the frequency-current relationship of single neurons (a step function in this model [23]). Note that the probability Ψa(ρe, ρi,(cid:104)n(cid:105)) is the same for both excitatory and inhibitory neurons because, in the network under consideration, excitatory and inhibitory neurons occupy topologically equivalent positions. Thus, Ψe(ρe, ρi,(cid:104)n(cid:105)) = Ψi(ρe, ρi,(cid:104)n(cid:105)) ≡ Ψ(ρe, ρi,(cid:104)n(cid:105)), where Ψ(ρe, ρi,(cid:104)n(cid:105)) =(cid:80) n,k,l≥0 Θ(nJn+kJe+lJi−Ω)G(n,(cid:104)n(cid:105)) × (A1) Pk(geρec)Pl(giρic). Here, c = cντ and c is the mean degree. Θ(x) is the Heaviside step function. Pq(c) is the Poisson distribution function, Pq(c) = cqe−c/q!, (A2) and G(n,(cid:104)n(cid:105)) is the Gaussian distribution function, . G(n,(cid:104)n(cid:105)) = G0e−(n−(cid:104)n(cid:105))2/2σ2 and G0 is the normalization constant,(cid:80)∞ (A3) G(n,(cid:104)n(cid:105)) defines the probability that a neuron receives n spikes from endogenous stimulation per integration time τ . (cid:104)n(cid:105) is the mean number of spikes, σ2 is the variance, n=0 G(n,(cid:104)n(cid:105)) = 1. We use (cid:104)n(cid:105) as the control parameter characterizing the endogenous stimulation. Note that Eqs. (1) and (A1) are asymptotically exact in the thermodynamic limit, N → ∞ [21, 23]. In numerical simulations, we use the algorithm ex- plained in [21]. We used the following model parameters (except when mentioned otherwise): N = 104, c = 103, Ω = 30, τ ν = 1, µeτ = 0.1, α = 0.7, and gi = 0.25. Throughout this paper we used 1/µe ≡ 1 as time unit and Je ≡ 1 as input unit. Following [55], we chose Ji = −3Je. We also used Jn = Je and σ2 = 10 for the amplitude and variance of the endogenous stimulation. These parameter choices have been discussed in [21 -- 23]. Appendix B: The activation threshold of ILS In this appendix we show how the activation thresh- old Ath of ILS depends on the endogenous stimula- tion (cid:104)n(cid:105) near the SNIC bifurcation. Since we consider 8 Ψe = Ψi ≡ Ψ, this implies that there is only one steady state equation, ρ = Ψ(ρ,(cid:104)n(cid:105)). In this case, the SNIC bifurcation occurs when (ρ,(cid:104)n(cid:105)) = 1, ∂Ψ ∂ρ (B1) (B2) which determines the critical point nc2. In the region of ILS, close to the bifurcation, at 0 < nc2 − (cid:104)n(cid:105) (cid:28) nc2, we can study low activity fluctuations δρ((cid:104)n(cid:105)) = ρ((cid:104)n(cid:105)) − ρ(nc2) near the SNIC bifurcation. To find how the activity fluctuations depend on the stimulation (cid:104)n(cid:105), we apply the Taylor expansion to Ψ(ρ,(cid:104)n(cid:105)) in Eq. (B1) over δρ((cid:104)n(cid:105)) and δn = (cid:104)n(cid:105) − nc2 up to the second order in δρ((cid:104)n(cid:105)), δρ((cid:104)n(cid:105)) ≈ ∂Ψ ∂(cid:104)n(cid:105) δn + ∂Ψ ∂ρ + ∂2Ψ ∂ρ2 (δρ)2, (B3) where the derivatives of Ψ are taken at nc2. Using Eq. (B2), we find where ρ((cid:104)n(cid:105)) − ρ(nc2) ≈ −C(cid:112)nc2 − (cid:104)n(cid:105), (cid:17)−1 (cid:16) ∂2Ψ (cid:115) C = −2 ∂Ψ ∂(cid:104)n(cid:105) . ∂ρ2 Consequently, near the critical point nc2, the activation threshold Ath((cid:104)n(cid:105)) also follows Ath((cid:104)n(cid:105)) ∝(cid:112)nc2 − (cid:104)n(cid:105). (B4) (B5) (B6) [1] P. N. Banerjee, D. Filippi, and W. A. Hauser, Epilepsy [10] W. C. Stacey and B. Litt, Nat. Clin. Pract. Neurol. 4, Res. 85, 31 (2009). 190 (2008). [2] P. Ortinski and K. J. Meador, Epilepsy Behav. 5, 60 [11] V. K. Jirsa, W. C. Stacey, P. P. Quilichini, A. I. Ivanov, (2004). and C. Bernard, Brain 137, 2210 (2014). [3] P. Kwan and M. J. Brodie, N. Engl. J. Med. 342, 314 [12] P. Perucca, F. Dubeau, and J. Gotman, Brain 137, 183 (2000). (2014). [4] J. de Tisi, G. S. Bell, J. L. Peacock, A. W. McEvoy, W. F. Harkness, J. W. Sander, and J. S. Duncan, Lancet 378, 1388 (2011). [5] I. Najm, L. Jehi, A. Palmini, J. Gonzalez-Martinez, E. Paglioli, and W. Bingaman, Epilepsia 54, 772 (2013). [6] M. Ch´avez, M. Le Van Quyen, V. Navarro, M. Baulac, and J. Martinerie, IEEE Trans. Biomed. Eng 50, 571 (2003). [7] M. Le Van Quyen, J. Soss, V. Navarro, R. Robertson, M. Chavez, M. Baulac, and J. Martinerie, Clin. Neuro- physiol. 116, 559 (2005). [8] F. Mormann, R. G. Andrzejak, C. E. Elger, and K. Lehn- ertz, Brain 130, 314 (2007). [9] H. M. De Boer, M. Mula, and J. W. Sander, Epilepsy Behav. 12, 540 (2008). [13] Y. H. Raol and A. R. Brooks-Kayal, Prog. Mol. Biol. Transl. Sci. 105, 57 (2011). [14] G. Huberfeld, L. M. De La Prida, J. Pallud, I. Cohen, M. Le Van Quyen, C. Adam, S. Clemenceau, M. Baulac, and R. Miles, Nat. Neurosci. 14, 627 (2011). [15] M. A. Kramer, W. Truccolo, U. T. Eden, K. Q. Lepage, L. R. Hochberg, E. N. Eskandar, J. R. Madsen, J. W. Lee, A. Maheshwari, E. Halgren, et al., Proc. Natl. Acad. Sci. U.S.A. 109, 21116 (2012). [16] S. H. Strogatz, Nonlinear Dynamics And Chaos: With Applications To Physics, Biology, Chemistry, And Engi- neering (Perseus Books Group, New York, 1994). [17] M. Scheffer, J. Bascompte, W. A. Brock, V. Brovkin, S. R. Carpenter, V. Dakos, H. Held, E. H. Van Nes, M. Rietkerk, and G. Sugihara, Nature 461, 53 (2009). 9 [18] M. Breakspear, J. Roberts, J. R. Terry, S. Rodrigues, N. Mahant, and P. Robinson, Cereb. Cortex 16, 1296 (2006). [37] M. Goodfellow and P. Glendinning, J. Comput. Neurosci. 3, 1 (2013). [38] A. Rothkegel and K. Lehnertz, Europhys. Lett. 95, 38001 [19] P. Suffczynski, S. Kalitzin, and F. Lopes Da Silva, Neu- (2011). roscience 126, 467 (2004). [39] A. Rothkegel and K. Lehnertz, New J. Phys. 16, 055006 [20] D. Steyn-Ross and M. Steyn-Ross, Modeling Phase Tran- (2014). sitions in the Brain (Springer, New York, 2010). [21] K.-E. Lee, M. A. Lopes, J. F. F. Mendes, and A. V. Goltsev, Phys. Rev. E 89, 012701 (2014). [22] M. Lopes, K.-E. Lee, A. Goltsev, and J. Mendes, Phys. Rev. E 90, 052709 (2014). [40] R. Boehringer, D. Polygalov, A. J. Huang, S. J. Mid- dleton, V. Robert, M. E. Wintzer, R. A. Piskorowski, V. Chevaleyre, and T. J. McHugh, Neuron 94, 642 (2017). [41] M. P. Richardson, J. Neurol. Neurosurg. Psychiatr. 83, [23] A. V. Goltsev, F. V. de Abreu, S. N. Dorogovtsev, and 1238 (2012). J. F. F. Mendes, Phys. Rev. E 81, 061921 (2010). [42] F. Wendling, F. Bartolomei, J. Bellanger, and P. Chau- [24] S. N. Dorogovtsev and J. F. F. Mendes, Adv. Phys. 51, vel, Eur. J. Neurosci. 15, 1499 (2002). 1079 (2002). [25] M. E. J. Newman, SIAM Rev. 45, 167 (2003). [26] S. Boccaletti, V. Latora, Y. Moreno, M. Chavez, and [43] F. Wendling, P. Benquet, F. Bartolomei, and V. Jirsa, J. Neurosci. Methods 260, 233 (2016). [44] D. Holstein, A. Goltsev, and J. Mendes, Phys. Rev. E D.-U. Hwang, Phys. Rep. 424, 175 (2006). 87, 032717 (2013). [27] E. M. Izhikevich and G. M. Edelman, Proc. Natl. Acad. [45] P. Milanowski and P. Suffczynski, Int. J. Neural Syst. 26, Sci. U.S.A. 105, 3593 (2008). 1650053 (2016). [28] R. A. Stefanescu, R. Shivakeshavan, and S. S. Talathi, [46] F. L. da Silva and G. F. Harding, Epilepsy Res. 97, 278 Seizure 21, 748 (2012). (2011). [29] E. M. Izhikevich, Int. J. Bif. and Chaos 10, 1171 (2000). [30] P. J. Karoly, D. R. Freestone, R. Boston, D. B. Gray- den, D. Himes, K. Leyde, U. Seneviratne, S. Berkovic, T. OBrien, and M. J. Cook, Brain , aww019 (2016). [31] O. N. Markand, Semin. Neurol. 23, 7 (2003). [32] J. Rinzel and G. B. Ermentrout (MIT Press, Cambridge, USA, 1989) Chap. Analysis of neural excitability and os- cillations, pp. 135 -- 169. [47] S. N. Kalitzin, D. N. Velis, and F. H. L. da Silva, Epilepsy Behav. 17, 310 (2010). [48] T. Inouye, H. Sakamoto, K. Shinosaki, S. Toi, and S. Ukai, Electroencephalogr. Clin. Neurophysiol. 76, 205 (1990). [49] D. Gupta, P. Ossenblok, and G. van Luijtelaar, Med. Biol. Eng. Comput. 49, 555 (2011). [50] E. Sitnikova and G. van Luijtelaar, Epilepsy Res. 84, 159 [33] F. L. da Silva, W. Blanes, S. N. Kalitzin, J. Parra, P. Suf- (2009). fczynski, and D. N. Velis, Epilepsia 44, 72 (2003). [51] G. Van Luijtelaar, A. Hramov, E. Sitnikova, and A. Ko- [34] Y. Wang, A. J. Trevelyan, A. Valentin, G. Alarcon, P. N. Taylor, and M. Kaiser, PLOS Comput. Biol. 13, e1005475 (2017). ronovskii, Clin. Neurophysiol. 122, 687 (2011). [52] O. Sporns, D. R. Chialvo, M. Kaiser, and C. C. Hilgetag, Trends Cogn. Sci. 8, 418 (2004). [35] F. Wendling, F. Bartolomei, J.-J. Bellanger, J. Bourien, [53] Z. F. Mainen and T. J. Sejnowski, Science 268, 1503 and P. Chauvel, Brain 126, 1449 (2003). (1995). [36] M. Goodfellow, K. Schindler, and G. Baier, Neuroimage [54] E. Schneidman, B. Freedman, and I. Segev, Neural Com- 55, 920 (2011). put. 10, 1679 (1998). [55] D. J. Amit and N. Brunel, Cereb. Cortex 7, 237 (1997).
1607.06251
2
1607
2016-11-07T14:46:30
Next generation neural mass models
[ "q-bio.NC" ]
Neural mass models have been actively used since the 1970s to model the coarse grained activity of large populations of neurons and synapses. They have proven especially useful in understanding brain rhythms. However, although motivated by neurobiological considerations they are phenomenological in nature, and cannot hope to recreate some of the rich repertoire of responses seen in real neuronal tissue. In this chapter we consider the $\theta$-neuron model that has recently been shown to admit to an exact mean-field description for instantaneous pulsatile interactions. We show that the inclusion of a more realistic synapse model leads to a mean-field model that has many of the features of a neural mass model coupled to a further dynamical equation that describes the evolution of network synchrony. A bifurcation analysis is used to uncover the primary mechanism for generating oscillations at the single and two population level. Numerical simulations also show that the phenomena of event related synchronisation and desynchronisation are easily realised. Importantly unlike its phenomenological counterpart this \textit{next generation neural mass model} is an exact macroscopic description of an underlying microscopic spiking neurodynamics, and is a natural candidate for use in future large scale human brain simulations.
q-bio.NC
q-bio
Next generation neural mass models Stephen Coombes and ´Aine Byrne Centre for Mathematical Medicine and Biology, School of Mathematical Sciences, University of Nottingham, University Park, Nottingham, NG7 2RD, UK. Summary. Neural mass models have been actively used since the 1970s to model the coarse grained activity of large populations of neurons and synapses. They have proven especially useful in understanding brain rhythms. However, although motivated by neurobiological con- siderations they are phenomenological in nature, and cannot hope to recreate some of the rich repertoire of responses seen in real neuronal tissue. In this chapter we consider the θ-neuron model that has recently been shown to admit to an exact mean-field description for instan- taneous pulsatile interactions. We show that the inclusion of a more realistic synapse model leads to a mean-field model that has many of the features of a neural mass model coupled to a further dynamical equation that describes the evolution of network synchrony. A bifurcation analysis is used to uncover the primary mechanism for generating oscillations at the sin- gle and two population level. Numerical simulations also show that the phenomena of event related synchronisation and desynchronisation are easily realised. Importantly unlike its phe- nomenological counterpart this next generation neural mass model is an exact macroscopic description of an underlying microscopic spiking neurodynamics, and is a natural candidate for use in future large scale human brain simulations.1 1 Introduction The term neural mass model is often used to refer to low dimensional models that aim to describe the coarse grained activity of large populations of neurons and synapses. They are typically cast as systems of ordinary differential equations (ODEs) and in their modern in- carnations are exemplified by variants of the two dimensional Wilson-Cowan model [1]. This model tracks the activity of an excitatory population of neurons coupled to an inhibitory population. With the augmentation of such models by more realistic forms of synaptic and network interaction they have proved especially successful in providing fits to neuroimaging data. Historically one of the first examples in this area is the Zetterberg model [2] for the electroencephalogram (EEG) rhythm. This is based on previous ideas developed by Lopes da Silva and colleagues [3, 4] and is built from three interacting neural mass models, as a minimal model of a cortical column. The first represents a population of pyramidal cells, the second a population of excitatory interneurons, and the third a population of inhibitory in- terneurons. Since its introduction the Zetterberg model has become more widely popularised by the work of Jansen and Rit [5] and used to explain epileptic brain dynamics, particu- larly by Wendling and colleagues, as recently reviewed in [6]. Another well known neural mass model is that of Liley, which pays particular attention to the role of synaptic reversal 1Contribution to the Workshop "Nonlinear Dynamics in Computational Neuroscience: from Physics and Biology to ICT" held in Turin (Italy) in September 2015. 6 1 0 2 v o N 7 ] . C N o i b - q [ 2 v 1 5 2 6 0 . 7 0 6 1 : v i X r a 2 Stephen Coombes and ´Aine Byrne potentials, and see [7] for a discussion of this model within the context of Freeman's ideas on the importance of chaos for cognition and perception [8]. As well as proving useful for understanding EEG rhythms ranging from delta (1− 4 Hz) through to gamma (30− 70 Hz) [9], neural mass models have been used to describe brain resonance phenomena [10], resting brain state activity [11] and are very popular in the neuroimaging community. In this latter in- stance they are often used for model driven fusion of multiple neuroimaging modalities, such as EEG and functional magnetic resonance imaging (fMRI) [12], as well as to augment the dynamic causal modelling framework for understanding how event-related responses result from the dynamics of coupled neural populations [13]. Moreover, they are now an integral part of the Virtual Brain project that aims to deliver the first open simulation of the human brain based on individual large-scale connectivity [14], as well as play a key role in the neuro- computational modelling of neurological and psychiatric disorders [15]. This latter work is especially viable since neural mass models can incorporate accurate descriptions of synaptic processing, typically in the form of a synaptic response function that is driven by firing rate rather than by the arrival times of individual action potentials. However, it is important to remember that at heart all neural mass models to date are essentially phenomenological, with state variables that track coarse grained measures of the average membrane potential, popu- lation firing rate or synaptic activity. At best they are expected to provide appropriate levels of description for many thousands of near identical interconnected neurons with a preference to operate in synchrony. This latter assumption is especially important for the generation of a sufficiently strong physiological signal that can be detected non-invasively. The variation of synchrony within a neuronal population is believed to underly the decrease or increase of power seen in given EEG frequency bands. The former phenomenon is called event-related desynchronisation (ERD), and the latter event-related synchronisation (ERS) [16]. Unfortu- nately the assumption of synchrony within neural mass models means that they cannot hope to describe ERD and ERS, at least not at the single population level. Rather this sets a natural challenge for the next generation of neural mass models. It is precisely this issue that we take up in this chapter. As a starting point to move beyond the current neural mass models we draw inspiration from the physics of self-organised networks. In particular the observation of macroscopic co- herent states in large networks of coupled spiking neuron models has inspired a search for equivalent low-dimensional dynamical descriptions, and see [17] for a recent review of os- cillatory network dynamics in neuroscience. However, although the mathematical step from microscopic to macroscopic dynamics has proved elusive for the majority of spiking mod- els the θ-neuron model has proven amenable to such a reduction for pulsatile coupling by Luke et al. [18]. A similar approach by Montbri´o et al. [19] has been used to reduce net- works of quadratic integrate-and-fire neurons. Here we show how to naturally augment these approaches to incorporate the biologically realistic forms of synaptic coupling that are com- monly adopted within current neural mass models. In this way we arrive at the first instance of a next generation neural mass model, with a derived (as opposed to postulated) population firing rate that couples directly to a dynamical variable describing population synchrony. In §2 we discuss the main elements of synaptic processing that are incorporated within standard neural mass models, and give a heuristic description of how to close the equations of motion in terms of population firing rates. The same model of a synapse is used in §3, though this time driven by the spike times arising in a network of θ-neurons. For a large globally coupled network an exact mean field description is derived, and the form of the equations compared and contrasted with standard phenomenological neural mass models. In §4 we present a bifur- cation analysis for the single and two population mean-field models, and use this to highlight Next generation neural mass models 3 the primary mechanisms for generating population oscillations. Importantly we also show, through direct numerical simulations, that the model supports ERD and ERS. Finally in §5 we reflect upon the use of such models in future large scale human brain simulations, as well as their subsequent mathematical analysis. 2 Neural mass modelling Neural mass models generate brain rhythms using the notion of population firing rates, aiming to side-step the need for large scale simulations of more realistic networks of spiking neu- rons. However, both approaches often make use of the same level of description for synaptic processing, in a manner that we shall now clarify. At a synapse, presynaptic firing results in the release of neurotransmitters that causes a change in the membrane conductance of the post-synaptic neuron. This post-synaptic current may be written I = g(vsyn − v), where v is the voltage of the post-synaptic neuron, vsyn is its membrane reversal potential and g is a conductance. This conductance is proportional to the probability that a synaptic receptor channel is in an open conducting state. This probability depends on the presence and concentration of neurotransmitter released by the presynaptic neuron. The sign of vsyn relative to the resting potential (assumed to be zero) determines whether the synapse is excitatory (vsyn > 0) or inhibitory (vsyn < 0). The effect of some synapses can be described with a function that fits the shape of the post-synaptic response due to the arrival of action potential at the pre-synaptic release site. A post-synaptic conductance change g(t) would then be given by g(t) = ks(t − T ) for t ≥ T , where T is the arrival time of a pre-synaptic action potential, s(t) fits the shape of a realistic post-synaptic conductance, and k is a constant. A common (normalised) choice for s(t) is the α-function: s(t) = α2te−αtΘ (t), (1) where Θ is a Heaviside step function. The conductance change arising from a train of action potentials, with firing times T m, is given by g(t) = k∑ m s(t − T m). (2) If s is the Green's function of a linear differential operator, so that Qs = δ , then we may write (2) in the equivalent form This is indeed the case for the choice (1) for which δ (t − T m). m Qg = k∑ (cid:18) Q = 1 + (cid:19)2 . 1 α d dt (3) (4) In many neural population models it is assumed that the interactions are mediated by fir- ing rates rather than action potentials (spikes) per se. To see how this might arise we perform a short-time average of (3) over some time-scale τ and assume that s is sufficiently slow so that (cid:104)Qg(cid:105)t is approximately constant, where 1 (cid:104)x(cid:105)t = τ x(t(cid:48))dt(cid:48), (cid:90) t (5) t−τ 4 Stephen Coombes and ´Aine Byrne then we have that Qg = k f , where f is the instantaneous firing rate (number of spikes per time ∆). For a single neuron (real or synthetic) experiencing a constant drive it is natural to assume that this firing rate is a function of the drive alone. If for the moment we assume that a neuron spends most of its time close to rest such that vsyn − v ≈ vsyn, and absorb a factor vsyn into k, then for synaptically interacting neurons this drive is directly proportional to the conductance state of the presynaptic neuron. Thus for a single population of identically and globally coupled neurons operating synchronously we are led naturally to equations like: Qg = κ f (g), (6) for some strength of coupling κ. A common choice for the population firing rate function is the sigmoid f (g) = , (7) f0 1 + e−r(g−g0) which saturates to f0 for large g. This functional form, with threshold g0 and steepness param- eter r, is not derived from a biophysical model, rather it is seen as a physiologically consistent choice. The extension to multiple interacting populations is straight forward, and the popular Jansen-Rit model [5], provides a classic example of such a generalisation. This can be written in the form QEgP = κP f (gE − gI), QEgE = κE f (w1gP) + A, QIgI = κI f (w2gP), (8) which describes a network of interacting pyramidal neurons (P), inhibitory interneurons (I) and excitatory interneurons (E). Here, Qa is given by (4) under the replacement α → αa for a ∈ {E,I}, w1,2, κE,I,P are constants, and A is an external input. It has been used to model both normal and epileptic patterns of cortical activity and its bifurcation structure has been systematically analysed in [20, 21]. Despite its usefulness in describing certain large scale brain rhythms, and especially alpha (8− 13 Hz), it suffers the same deficiencies as all other neural mass models, namely it cannot track the level of synchrony within a neuronal population. 3 θ-neuron network and reduction The θ-neuron model or Ermentrout-Kopell canonical model is now widely known throughout computational neuroscience as a parsimonious model for capturing the firing and response properties of a cortical cell [22]. It is described by a purely one dimensional dynamical system evolving on a circle according to θ = (1− cosθ ) + (1 + cosθ )η, θ ∈ [−π,π), (9) d dt where η represents a constant drive. For η < 0 the θ-neuron supports a pair of equilibria θ±, with θ+ < 0 and θ− > 0, and no equilibria for η > 0. In the former case the equilibria at θ+ is stable and the one at θ− unstable. In neurophysiological terms, the unstable fixed point at θ− is a threshold for the neuron model. Any initial conditions with θ ∈ (θ+,θ−) will be attracted to the stable equilibrium, while initial data with θ > θ− will make a large excursion around the circle before returning to the rest state. For η > 0 the θ-neuron oscillates with √ η. When η = 0 the θ-neuron is poised at a saddle-node on an invariant circle frequency 2 (SNIC) bifurcation. Next generation neural mass models 5 A network of θ-neurons can be described with the introduction of an index i = 1, . . . ,N and the replacement η → ηi + Ii, where Ii describes the synaptic input current to neuron i. For a globally coupled network this can be written in the form Ii = g(t)(vsyn − vi) for some global conductance g and local voltage vi. As a model for the conductance we take the form used in (3) and write Qg(t) = k N δ (t − T m j ), ∑ m∈Z (10) N ∑ j=1 where T m is the mth firing time of the jth neuron. These are defined to happen every time j θ j increases through π. It is well known that the θ-neuron model is formally equivalent to a quadratic integrate-and-fire model for voltage dynamics [23] under the transformation vi = tan(θi/2) (so that cosθi = (1 − v2 i )). This voltage relationship allows us to write the network dynamics as i ) and sinθi = 2vi/(1 + v2 i )/(1 + v2 θi = (1− cosθi) + (1 + cosθi)(ηi + g(t)vsyn)− g(t)sinθi, d dt Qg = 2 k N N ∑ j=1 P(θ j). (11) (12) j ) = δ (θ j(t)− π) θ j(T m Here P(θ ) = δ (θ −π) and is periodically extended such that P(θ ) = P(θ +2π), and we have used the result that δ (t − T m j ). For the case that vsyn (cid:29) vi, Q = 1, and P(θ ) has a shape of the form (1− cos(θ ))n for some positive integer n we recover the model of Luke et al. [18]. In this case these authors have shown how to obtain an exact mean- field reduction making use of the Ott-Antonsen (OA) ansatz. The same style of reduction has also been used by Paz´o and Montbri´o to study pulse-coupled Winfree networks [24]. The OA anstaz was originally used to find solutions on a reduced invariant manifold of the Ku- ramoto model [25], and essentially assumes that the distribution of phases as N → ∞ has a simple unimodal shape, capable of describing synchronous (peaked) and asynchronous (flat) distributions. In the following we show how their reduction approach extends to the more bio- logically realistic model described by (11)-(12) that includes synaptic reversal potentials and causal non-instantaneous synaptic responses. We note that even in the limit of fast synaptic interactions we do not recover models of the type described in [18, 24, 19] due to our focus on conductance changes and the inclusion of voltage shunts. 3.1 Mean field reduction In the following we shall choose the background drives ηi to be random variables drawn from a Lorentzian distribution L(η) with L(η) = 1 π ∆ (η − η0)2 + ∆ 2 , (13) where η0 is the centre of the distribution and ∆ the width at half maximum. For the choice of Q we shall take equation (4). In the coupled network, and if the frequencies of the individual neurons are similar enough, then one may expect some degree of phase locking (ranging from synchrony to asynchrony), itself controlled in part by the the time-to-peak, 1/α, of the synaptic filter. In the limit N → ∞ the state of the network at time t can be described by a continuous probability distribution function ρ(η,θ ,t), which satisfies the continuity equation (arising from the conservation of oscillators): 6 Stephen Coombes and ´Aine Byrne ∂ ∂t ρ + ∂ ∂θ ρc = 0, c = lim N→∞ 1 N N ∑ j=1 d dt θ j. (14) The global drive to the network given by the right hand side of (12) can be constructed as lim N→∞ 1 N N ∑ j=1 P(θ j) = Hence, (cid:90) 2π 0 (cid:90) ∞ −∞ dθ dρ(η,θ ,t)P(θ ). c = (1− cosθ ) + (1 + cosθ )(η + gvsyn)− gsinθ , (cid:90) 2π 0 (cid:90) ∞ −∞ Qg = k π ∑ m∈Z dθ dρ(η,θ ,t)eim(θ−π), (15) (16) (17) (20) (21) where we have used the result that 2πP(θ ) = ∑m∈Z eim(θ−π). The formula for c above may be written conveniently in terms of e±iθ as c = f eiθ + h + f e−iθ , (18) where f = ((η − 1) + vsyng + ig)/2 and h = (η + 1) + vsyng, and f denotes the complex conjugate of f . The OA ansatz assumes that ρ(η,θ ,t) has the product structure ρ(η,θ ,t) = L(η)F(η,θ ,t). Since F(η,θ ,t) should be 2π periodic in θ it can be written as a Fourier series: (cid:41) F(η,θ ,t) = 1 2π ∞ ∑ n=1 1 + Fn(η,t)einθ + cc , (19) where cc denotes complex conjugate. The insight in [26] was to restrict the Fourier coef- ficients such that Fn(η,t) = a(η,t)n, where a(η,t) ≤ 1 to avoid divergence of the series. There is also a further requirement that a(η,t) can be analytically continued from real η into the complex η-plane, and that this continuation has no singularities in the lower half η-plane, and that a(η,t) → 0 as Imη → −∞. If we now substitute (18) into the continuity equation (14), use the OA ansatz, and balance terms in eiθ we obtain an evolution equation for a(η,t) as It is now convenient to introduce the Kuramoto order parameter ∂ ∂t a− ia2 f − iah− i f = 0. (cid:90) 2π (cid:90) ∞ Z(t) = dθ −∞ 0 dηρ(η,θ ,t)eiθ , (cid:82) 2π where Z ≤ 1. Using the OA ansatz (and using the orthogonality properties of eiθ , namely 0 eipθ eiqθ dθ = 2πδp+q,0) we then find that Z(t) = dηL(η)a(η,t). (22) By noting that the Lorentzian (13) has simple poles at η± = η0 ± i∆ the integral in (22) may be performed by choosing a large semi-circle contour in the lower half η-plane. This yields Z(t) = a(η−,t), giving Z(t) = a(η+,t). Hence, the dynamics for g given by (17) can be written as Qg = κ f (Z), where κ = k/π and −∞ (cid:40) (cid:90) ∞ f (Z) = ∑ m∈Z (−Z)m = Next generation neural mass models 1−Z2 Z < 1. 1 + Z + Z +Z2 , 7 (23) The dynamics for Z is obtained from (20) as dZ/dt = F (Z;η0,∆ ) + G (Z,g;vsyn), where F (Z;η0,∆ ) = −i (Z − 1)2 2 G (Z,g;vsyn) = i (Z + 1)2 + (Z + 1)2 vsyng− Z2 − 1 2 [−∆ + iη0] , 2 g. 2 (24) (25) Here we may view (24) as describing the intrinsic population dynamics and (25) the dynamics induced by synaptic coupling. Thus the form of the mean field model is precisely that of a neural mass model as given by equation (6). Importantly the firing rate f is a derived quantity that is a real function of the complex Kuramoto order parameter for synchrony. This in turn is described by a complex ODE with parameters from the underlying microscopic model. Fig. 1. Comparison between the reduced mean field model (blue) and a simulation of a net- work of 500 θ-neurons (red). Left: Phase plane projection for the Kuramoto order parame- ter Z = ReiΨ . Right: Phase plane projection for the synaptic conductance g. Here η0 = 20, ∆ = 0.5, vsyn = −10 , κ = 1, α = 0.95. To illustrate the validity of the reduction presented above we show a simulation of a network with N = 500 θ-neurons and the mean field model in Fig. 1. Here we plot the real and imaginary parts of Z which we write in the form ReiΨ for the mean field reduction and calculate as Z = N−1 ∑N j=1 eiθ j for the finite size network simulation. It is abundantly clear that the two realisations agree very well. If the size of the population in the large scale simulations is reduced then one can more easily see finite size fluctuations as expected. 4 Next generation neural mass model: analysis The mean field model derived in §3.1 is a natural candidate for a next generation neural mass model. It generalises the form of the phenomenological neural mass model whilst maintaining 8 Stephen Coombes and ´Aine Byrne contact with biological reality in that it preserves the notion of both population rate and syn- chrony. An almost identical model has recently been discussed by Laing [27], although here the focus was on a first order synapse model (namely Q = (1 + α−1d/dt)) with no provision for synaptic reversal potentials. In mathematical terms we are now faced with understanding the dynamics of a coupled system of ODEs given by Qg = κ f (Z), d dt Z = F (Z;η0,∆ ) + G (Z,g;vsyn), (26) with f , F and G given by (23), (24) and (25) respectively, and Q a linear differential operator such as given by (4). One practical way to assess the emergent behaviour of the model (26) under parameter variation is through numerical bifurcation analysis. We now pursue this for (26), as well as for its natural extension to cover two interacting populations. 4.1 Bifurcation diagrams We first consider the case of a purely inhibitory population. Using XPPAUT [28] we find that for a wide range of system parameters it is possible to find a Hopf bifurcation of a steady state to a periodic orbit under parameter variation of η0 (controlling the mean value of the background drive). To illustrate the relatively large volume of parameter space that can support oscillations by this mechanism we show a two parameter continuation of the Hopf bifurcation in (∆ ,η0) (controlling the shape of the distribution (13)) in Fig. 2, for several values of the coupling strength κ and reversal potential vsyn. Since the OA ansatz does not hold for ∆ = 0 (so that some degree of heterogeneity must be present) and oscillations are only possible for η0 > 0, the bifurcation diagrams are only presented for the scenario ∆ ,η0 > 0. Fig. 2. Two parameter continuation of a Hopf bifurcation in the single population model de- scribed by (26) using (4) with α = 1. Left: Curves obtained for various values of κ with vsyn = −5. Right: Curves obtained under for various values of vsyn with κ = 1. In both dia- grams the area under the curves represents the parameter window for oscillatory behaviour. Next generation neural mass models 9 Suppose now that we have two populations, one excitatory and one inhibitory, with re- ciprocal connections. Introducing the labels E and I for each population then the natural generalisation of (26) is Qabgab = κab f (Zb), d dt Za = Fa(Za) +∑ b Gb(Za,gab), (27) where a,b ∈ {E,I}. Here, Qab is obtained from (4) under the replacement α → αab, Fa(Za) = F (Za;ηa syn). The system of equations (27) 0 ,∆ a) and Gb(Za,gab) = G (Za,gab;vab Fig. 3. Bifurcation diagram for the reciprocally connected PING network defined by (27) under variation of κIE, for both the excitatory (blue) and inhibitory (red) populations. Solid lines: stable; dashed lines: unstable. Circles show maximum and minimum values of f (ZE ) and f (ZI) over one period of oscillation when no steady states are stable. The inset shows syn = −10, a PING rhythm for κIE = 0.65. Parameters: αEI = 0.8, αIE = 10, κEI = 0.5, vEI vIE syn = 10, ηE 0 = 0, ∆ E = ∆ I = 0.5, κEE = κII = 0. 0 = 10, ηI generalises those recently presented by Laing [27] (to include reversal potentials, higher or- der synapses and self coupling), who highlighted the ability of such networks to produce a so-called pyramidal-interneuronal network gamma (PING) rhythm [29], as shown in the inset of Fig. 3. Here we also show a bifurcation digram as a function of κIE, when κEE = κII = 0, which shows that periodic behaviour can be destroyed in a supercritical Hopf bifurcation as κIE is decreased. In Fig. 4 we show bifurcation diagrams under the variation of κEI and ηI 0. We see that when κEI is decreased periodic behaviour can be destroyed in a supercritical Hopf bifurcation. With an increase in ηI 0 it is also possible to generate a supercritical Hopf bifurcation to terminate the PING rhythm. The inclusion of self coupling leads to a wide variety of bifurcations, as seen in Fig. 5. As ηI 0 is increased we observe the appearance of oscillatory behaviour through a supercritical Hopf bifurcation, which is destroyed at a second supercritical Hopf bifurcation when ηI 0 is increased further. Note the appearance/disappearance of period doubling through two period 10 Stephen Coombes and ´Aine Byrne Fig. 4. Corresponding bifurcation diagrams to Fig. 3 under variation in κEI (left) and ηI 0 (right), for both the excitatory (blue) and inhibitory (red) populations. Note that PING rhythms can be terminated by either decreasing the strength of coupling to the excitatory pop- ulation from the inhibitory population or increasing the natural frequency of the inhibitory population. Parameters as in Fig. 3 with κIE = 0.9 for both. doubling bifurcations on this branch of periodic orbits. We also observe the appearance and disappearance of an isola of periodic orbits through two saddle node bifurcations of peri- odic orbits. The first saddle node occurs before the second Hopf bifurcation, i.e. there exists two stable periodic orbits for this window of parameter space. Further increasing ηI 0 leads to another saddle node bifurcation of periodic orbits, shortly followed by a torus bifurcation and then a saddle-node on invariant circle bifurcation which destroys the unstable branch of the periodic orbit. The stable branch of the periodic orbit is destroyed at a supercritical Hopf bifurcation, as ηI 0 is increased further. Along the unstable fixed point branch there are four Hopf bifurcations all of which either create or destroy unstable periodic behaviour. Be- tween the second and third of these bifurcations there are two torus bifurcations, one on each periodic orbit. 0 = −20 and ηI The inset in Fig. 5 shows the behaviour of the system for ηI 0 = 25 respec- tively. In both cases the excitatory population has two frequencies and the lower frequency is synchronised to the inhibitory population. For ηI 0 = 25, the system follows the orbit created by the isola, and there are two peaks in the (inhibitory) firing rate per period. 4.2 Event related synchronisation and desynchronisation system), where J(t) is a smoothed rectangular pulse of the form J(t) =(cid:82) t−∞ ηD(t − s)A(s)ds, Here we show that a time varying input to a single population model given by (26) can both disrupt and enhance the degree of synchrony within the population. We include such a drive under the replacement η0 → η0 + J(t) (describing a homogeneous drive to the microscopic for A(t) = σΘ (t − T )Θ (T + τ − t) and ηD(t) is an α-function of the form (1) under the re- placement α → αD. In Fig. 6 we show that for the inhibitory population considered in §3.1 operating in its oscillatory regime that such a drive can initially cause a population to desyn- chronise (during the pulse) though that upon termination of the pulse the system can rebound and generate a stronger spectral power peak than seen before the presentation of the pulse (before relaxing back to the undriven periodic orbit). Thus, in contrast to a standard neural mass model, the mean field model (26) is mechanistically able to support the phenomena Next generation neural mass models 11 Fig. 5. Bifurcation diagram for ηI 0 showing the exotic behaviour of the two population model when self coupling is reintroduced. Solid lines: stable; dashed lines: unstable; green (blue) dotted line: stable (unstable) oscillations; orange crosses: period doubling bifurcations; red stars: torus bifurcations. Of particular interest is the appearance/disappearance of an isola at 0 (cid:39) 15− 50. The insets shows the firing rate for the excitatory (blue) and inhibitory (red) ηI populations. In the upper inset ηI 0 = 25. Other parameters: αEE = 1, αEI = 0.7, αIE = 1.4, αII = 0.4, κEE = 1.5, κEI = 2, κIE = 1, κII = 3, vEE syn = 10, syn = −vEI syn = −12, ηE vIE 0 = −20 and in the lower inset ηI 0 = 20, ∆ E = ∆ I = 0.5. syn = 8, vII of ERD and ERS. A more thorough discussion of this observation can be found in Byrne et al. [30], where the model is used to describe magnetoencephalography (MEG) data for movement related beta decrease and post-movement beta rebound [16]. 5 Discussion The desire to understand large scale brain dynamics as observed using EEG, MEG and fMRI has prompted the increasing use of computational models [31]. Many of these approaches, such as exemplified by the Virtual Brain project [14], make use of networks of interconnected neural mass models. However, the inability of a single neural mass model to support the well documented phenomena of ERS and ERD reminds us that these phenomenological models could be improved upon. Of course, building more detailed biophysically realistic models of neurons and their interactions would improve this state of affairs, though at a price. This being not only computational complexity but our ability to interpret the behaviour of very high di- mensional models in a meaningful way. The model that we have presented here is very much in the original spirit of neural mass modelling, yet importantly it can be interpreted directly in terms of an underlying spiking model. Moreover the derived structure of the macroscopic equations can be viewed as a modification of the standard neural mass framework whereby the firing rate of the system is now coupled to the degree of synchrony within a population. Given the recent success of this model in explaining beta rebound [30], we advocate strongly for its subsequent use in future population-level modelling approaches for understanding in vivo brain activity states. Indeed we would like to think that this is a first example of a next 12 Stephen Coombes and ´Aine Byrne Fig. 6. Left: Phase plane for Z = ReiΨ , demonstrating the behaviour of the system (26) in response to a drive in the form of a smoothed rectangular pulse. The blue curve represents the system before the pulse arrives, as it settles to its non-perturbed dynamics (t < T ),the red curve demonstrates how the system behaves when the pulse is switched on (at t = T ) and the green how the system reacts once the drive is switched off (at t = T + τ). Right: The corresponding spectrogram of the synaptic current demonstrating rebound (an enhanced spectral peak on cessation of the applied pulse). Parameter values as in Fig. 1, with T = 40, τ = 12, η0 = 21.5, αD = 6 and σ = 15. generation neural mass model and that there will be others to follow. For phase oscillator sin- gle neuron models, this is intrinsically linked to the mathematical challenge of generalising the OA approach, whilst for more general conductance based neurons one might appeal to the McKean-Vlasov-Fokker-Planck approach of Baladron et al. [32]. However, even before ex- tending the model we have presented here to include more biological features (such as action potential generation, dendritic processing, and stochasticity) there is still much to be done in understanding the response of the model to input. Of particular utility would be an under- standing of the response to periodic forcing, as this would be a precursor to understanding patterns of phase-locking, clustering, chaos, and the multiplicity of attractors expected at the network level. Moreover, given that neural mass models are themselves the building blocks of neural field models [33] it would be interesting to pursue the analysis of bumps, waves and patterns in continuum versions of the model presented here, along the lines recently devel- oped by Laing for both synaptic and gap junction coupled systems [34, 27]. Acknowledgement. SC was supported by the European Commission through the FP7 Marie Curie Initial Training Network 289146, NETT: Neural Engineering Transformative Tech- nologies. References 1. H R Wilson and J D Cowan. Excitatory and inhibitory interactions in localized popula- tions of model neurons. Biophysical Journal, 12:1–24, 1972. Next generation neural mass models 13 2. L H Zetterberg, L Kristiansson, and K Mossberg. Performance of a model for a local neuron population. Biological Cybernetics, 31:15–26, 1978. 3. F H Lopes da Silva, A Hoeks, H Smits, and L H Zetterberg. Model of brain rhythmic activity: The alpha-rhythm of the thalamus. Kybernetik, 15:27–37, 1974. 4. F H Lopes da Silva, A van Rotterdam, P Barts, E van Heusden, and W Burr. Models of neuronal populations: the basic mechanisms of rhythmicity. Progress in Brain Research, 45:281–308, 1976. 5. B H Jansen and V G Rit. Electroencephalogram and visual evoked potential generation in a mathematical model of coupled cortical columns. Biological Cybernetics, 73:357–366, 1995. 6. F Wendling, P Benquet, F Bartolomei, and V Jirsa. Computational models of epilepti- form activity. Journal of Neuroscience Methods, 260:233–251, 2016. 7. M P Dafilis, F Frascoli, P J Cadusch, and D T J Liley. Chaos and generalised multista- bility in a mesoscopic model of the electroencephalogram. Physica D, 238:1056–1060, 2009. 8. W J Freeman. Tutorial on neurobiology: From single neurons to brain chaos. Interna- tional Journal of Bifurcation and Chaos, 2:451–482, 1992. 9. R C Sotero, N J Trujillo-Barreto, Y Iturria-Medina, F Carbonell, and J C Jimenez. Real- istically coupled neural mass models can generate EEG rhythms. Neural Computation, 19:478–512, 2007. 10. A Spiegler, T R Knosche, K Schwab, J Haueisen, and F M Atay. Modeling brain resonance phenomena using a neural mass model. PLoS Computational Biology, 7(12):e1002298, 2011. 11. G Deco, V K Jirsa, and A R McIntosh. Emerging concepts for the dynamical organization of resting-state activity in the brain. Nature Reviews Neuroscience, 12:43–56, 2011. 12. P Valdes-Sosa, J M Sanchez-Bornot, R C Sotero, Y Iturria-Medina, Y Aleman-Gomez, J Bosch-Bayard, F Carbonell, and T Ozaki. Model driven EEG/fMRI fusion of brain oscillations. Human Brain Mapping, 30:2701–21, 2009. 13. R Moran, D A Pinotsis, and K Friston. Neural masses and fields in dynamic causal modeling. Frontiers in Computational Neuroscience, 7(57):1–12, 2013. 14. P Sanz-Leon, S A Knock, A Spiegler, and V K Jirsa. Mathematical framework for large- scale brain network modeling in The Virtual Brain. NeuroImage, 111:385–430, 2015. 15. B S Bhattacharya and F N Chowdhury, editors. Validating Neuro-Computational Models of Neurological and Psychiatric Disorders. Springer, 2015. 16. G Pfurtscheller and F H Lopes da Silva. Event-related EEG/MEG synchronization and desynchronization: basic principles. Clinical Neurophysiology, 110:1842–1857, 1999. 17. P Ashwin, S Coombes, and R Nicks. Mathematical frameworks for oscillatory network dynamics in neuroscience. Journal of Mathematical Neuroscience, 6(2), 2016. 18. T B Luke, E Barreto, and P So. Complete classification of the macroscopic behaviour of a heterogeneous network of theta neurons. Neural Computation, 25:3207–3234, 2013. 19. E Montbri´o, D Paz´o, and A Roxin. Macroscopic description for networks of spiking neurons. Physical Review X, 5:021028, 2015. 20. A Spiegler, S J Kiebel, F M Atay, and T R Knosche. Bifurcation analysis of neural mass models: Impact of extrinsic inputs and dendritic time constants. NeuroImage, 52:1041– 1058, 2010. 21. J Touboul, F Wendling, P Chauvel, and O Faugeras. Neural mass activity, bifurcations, and epilepsy. Neural Computation, 23:3232–3286, 2011. 22. G B Ermentrout and N Kopell. Parabolic bursting in an excitable system coupled with a slow oscillation. SIAM Journal on Applied Mathematics, 46:233–253, 1986. 14 Stephen Coombes and ´Aine Byrne 23. P E Latham, B J Richmond, P G Nelson, and S Nirenberg. Intrinsic Dynamics in Neu- ronal Networks . I . Theory. Journal of Neurophysiology, 83:808–827, 2000. 24. D Paz´o and E Montbri´o. Low-dimensional dynamics of populations of pulse-coupled oscillators. Physical Review X, 4:011009, 2014. 25. Y Kuramoto. Collective synchronization of pulse-coupled oscillators and excitable units. Physica D, 50:15–30, 1991. 26. E Ott and T M Antonsen. Low dimensional behavior of large systems of globally coupled oscillators. Chaos, 18:037113, 2008. 27. C R Laing. Computational Models of Brain and Behavior, chapter Phase oscillator net- work models of brain dynamics. Wiley-Blackwell, 2016. 28. G B Ermentrout. Simulating, analyzing, and animating dynamical systems: A guide to XPPAUT for researchers and students. SIAM Books, Philadelphia, 2002. 29. C Borgers and N Kopell. Synchronization in networks of excitatory and inhibitory neu- 30. rons with sparse, random connectivity. Neural Computation, 15:509–538, 2003. ´A Byrne, M J Brookes, and S Coombes. A mean field model for movement induced changes in the β rhythm. NeuroImage, submitted, 2016. 31. I Bojak and M Breakspear. Neuroimaging, neural population models for. In Encyclope- dia of Computational Neuroscience, pages 1–29. Springer, 2014. 32. J Baladron, D Fasoli, O Faugeras, and J Touboul. Mean field description of and propaga- tion of chaos in networks of Hodgkin-Huxley and FitzHugh-Nagumo neurons. Journal of Mathematical Neuroscience, 2(10), 2012. 33. S Coombes, P beim Graben, R Potthast, and J J Wright, editors. Neural Field Theory. Springer Verlag, 2014. 34. C R Laing. Exact neural fields incorporating gap junctions. SIAM Journal on Applied Dynamical Systems, 14:1899–1929, 2015.
1610.01217
2
1610
2016-10-08T05:11:51
New hallmarks of criticality in recurrent neural networks
[ "q-bio.NC" ]
A rigorous understanding of brain dynamics and function requires a conceptual bridge between multiple levels of organization, including neural spiking and network-level population activity. Mounting evidence suggests that neural networks of cerebral cortex operate at criticality. How operating near this network state impacts the variability of neuronal spiking is largely unknown. Here we show in a computational model that two prevalent features of cortical single-neuron activity, irregular spiking and the decline of response variability at stimulus onset, are both emergent properties of a recurrent network operating near criticality. Importantly, our work reveals that the relation between the irregularity of spiking and the number of input connections to a neuron, i.e., the in-degree, is maximized at criticality. Our findings establish criticality as a unifying principle for the variability of single-neuron spiking and the collective behavior of recurrent circuits in cerebral cortex.
q-bio.NC
q-bio
New hallmarks of criticality in recurrent neural networks Yahya Karimipanah, Zhengyu Ma and Ralf Wessel Department of Physics, Washington University, St. Louis, Missouri 63130 A rigorous understanding of brain dynamics and function requires a conceptual bridge between including neural spiking and network-level population activity. multiple levels of organization, Mounting evidence suggests that neural networks of cerebral cortex operate at criticality. How operating near this network state impacts the variability of neuronal spiking is largely unknown. Here we show in a computational model that two prevalent features of cortical single-neuron activ- ity, irregular spiking and the decline of response variability at stimulus onset, are both emergent properties of a recurrent network operating near criticality. Importantly, our work reveals that the relation between the irregularity of spiking and the number of input connections to a neuron, i.e., the in-degree, is maximized at criticality. Our findings establish criticality as a unifying principle for the variability of single-neuron spiking and the collective behavior of recurrent circuits in cerebral cortex. I. Introduction II. Results Linking the evolutionary-derived multi-scale organiza- tion of the brain to neural dynamics and computation represents a major challenge in systems neuroscience [1]. Among the vast spectrum of spatial and temporal scales of brain activity, two experimentally accessible levels of brain organization are (i) the single-neuron spiking and (ii) the population activity of the network in which the neurons are embedded to various degrees. Single-neuron spiking in cerebral cortex is characterized by statistical properties, such as irregular spiking [2] and reduced vari- ability during sensory stimulation [3]. Population activ- ity is characterized by complex spatiotemporal activity, including scale-free activity [4], which is predicted to oc- cur for a network state near criticality [5]. These obser- vations at two adjacent levels of brain organization raise the question, to what extent the network state controls the variability of single-neuron spiking? Irregular spiking, defined as the mean coefficient of variation (CV) being larger than one, is known as one of the most widespread features of cortical activity in vivo [2, 6]. Theoretical studies have linked the prevalence of irregular spiking to the fluctuations of synaptic inputs at the sub-threshold regime [7, 8]. Despite the popularity of this hypothesis, here we provide an alternative scenario, by which the onset of irregular spiking can simply emerge at the transition between two phases of order and chaos. Further, using a computational model we show that even at the presence of other mechanisms for irregular spiking (CV > 1), neuronal activity shows maximum irregular- ity when the network resides at criticality. Moreover, in addition to irregular spiking, we show that criticality also gives rise to pronounced decline in neural variability after the onset of stimulus, as well as maximum correlation be- tween a number of single-neuron properties such as their CV's and firing rates. Our findings provide us with ro- bust measures of critical dynamics that at the same time could further our understanding about the implications of criticality for brain dynamics and function. To explore the impact of critical dynamics on the statistics of single-neuron spiking, we used a model net- work consisting of excitatory binary probabilistic neurons with sparse connectivity and external inputs (see fig.1a and Methods). This model permits the use of the maxi- mum eigenvalue λ of the transition probability matrix Pij as a control parameter to tune the network near the crit- ical point [9, 10]. Such tuning results in characteristic avalanche size distributions for the subcritical (λ < 1), critical (λ = 1), and supercritical (λ > 1) regime (fig.1 b-d). To investigate the hypothesized impact of the network state on the statistics of neuronal spiking, we simulated the network activity for different values of λ and quanti- fied the single-neuron spiking statistics using the CV of the inter-spike-interval distributions (fig.1 e). The CV is defined as the ratio of the standard deviation and the mean of the inter-spike-interval distribution for a given neuron. The irregular spiking is basically characterized by CV > 1, whereas CV = 1 is considered as Poisson spiking. It turned out that at the presence of a con- stant slow drive (see Methods) the CV values distributed around a mean greater than one, thus indicating irregular spiking. In contrast, small deviations of the network state towards either the subcritical or the supercritical regime resulted in CV values distributed around a mean of 1 or less. In summary, when tuning the network from the sub- critical to the supercritical state, the mean CV peaked near λ = 1 (fig.1f); with increasing network size, the peak moved closer towards λ = 1 and becomes narrower as well. This observation suggests that, at the large-size limit, the irregular spiking is an emergent property of re- current networks operating at the critical point (λ = 1). In other words, the scale-free fluctuations in network ac- tivity at criticality translate into irregular single-neuron spiking. The above results could be confounded by the finite network size. Importantly, the width of the mean CV as a function of λ decreased with increasing network size 2 Fig. 1: Irregular spiking emerges in a recurrent network operating at the critical point: (a) The model network consists of binary probabilistic model neurons with sparse connectivity (black) and weak external inputs (gray) to a fraction of the neurons. (b-d) Simulated spike trains (black raster), neuronal avalanches (gray), and corresponding avalanche size distributions (for 5 × 105 simulation time-steps) for a network of N = 500 neurons with 10% connectivity and external inputs η = 1/(10N ) to all neurons. Simulations were conducted for three different network states: subcritical ((b) λ = 0.9, blue), critical ((c) λ = 1.0, red), and supercritical ((d) λ = 1.1, green). (e) Inter-spike-interval CV distributions of simulated spike trains from a network of N = 5000 neurons with 3% connectivity and external inputs η = 1/(5N ) to all neurons. Simulations were conducted for the subcritical (λ = 0.9, blue), critical (λ = 1.02, red), and supercritical (λ = 1.06, green) network states. (f ) The average CV as a function of the maximum eigenvalue λ of the transition probability matrix Pij for three network sizes. Connectivity was 3% and external input was 1/(5N ) to all neurons. The curves were based on 13 values of λ within the range shown and were smoothed using Matlab spline. (g) The average population coupling and the change in mean response as a function of λ for a network of N = 5000 neurons with 3% connectivity and external inputs of strength η = 1/(5N ) applied to all neurons. To compute the change in mean response, we increased the external input strength by a factor of 10 half-way through the simulation, i.e., from η = 1/(5N ) to η = 2/N . and the peak location moved closer to λ = 1 (fig.1f). To further test whether the deviation of the peak from λ = 1 is indeed due to the finite size effect, we looked at two major characteristics of criticality, namely maximum correlations and mean response. In order to see how the average correlation among neurons behaves in terms of the control parameter λ, we computed a commonly-used measures of coordinated network activity that is known to be maximized at criticality. One such quantity is the average population coupling, which represents a measure of the overall level of correlated fluctuations within the network [11] (see Methods). Comparing the CV and the average population cou- pling reveals that they both peak at the same point (close to λ = 1), also similar to the change in mean response to an increase in external input. This demonstrates that the peak for CV coincides with the critical point, which is characterized by maximum average population coupling and change in mean response. Therefore, maximum CV could be regarded as a hallmark of critical dynamics for recurrent neural networks. The observed onset of irregular spiking near criticality (fig.1f) has an intuitive explanation in the extreme limits of connection strength. In the limit of weak connections (λ < 1), spiking is largely driven by the Poisson external input alone, thus resulting in Poisson like spiking. On the other hand, neurons become mostly active leading to more regular spiking in the limit of very strong connec- tions (λ > 1). However, it is at criticality (λ ≈ 1) that the scale-free fluctuations in network activity translate to single-neuron irregular spiking (hCV i > 1). Further- more, although there can be alternative mechanisms for irregular spiking, such as synchronous inputs (see discus- sion), we show that the impact of the network state on the statistics of single-neuron spiking is largely robust with respect to structured external inputs; the irregu- larity of spiking is always maximized near network crit- icality even at the presence of other sources of irregular spiking (fig.S1,fig.S2,fig.S3). The observed impact of the network state on the statistics of single-neuron spiking ( 3 Fig. 2: Solely for a recurrent network operating near criticality does the irregularity of spiking increase with a neuron's in-degree: (a, b) The inter-spike-interval CVs from simulated spike trains versus the neuron's in degree for a model network in three different states: (a) critical (λ = 1.02), (b) subcritical (λ = 0.9) and supercritical (λ = 1.1). Network parameters were N = 5000, 3% connectivity, and η = 1/(5N ) applied to all neurons. (c) Correlation between CV and in-degree as a function of λ. Other network parameters as in (a, b). The curves were smoothed using Matlab spline. The colored dots correspond to the distributions in (a, b). (d, e) CV versus rate for three network states. Network parameters as in (a, b). (f) Correlation between CV and rate as a function of λ. Other network parameters as in (a, b). (g, h) Population coupling versus in-degree for three network states. Network parameters as in (a, b). (i) Correlation between population coupling and in-degree as a function of λ. Other network parameters as in (a, b). fig.1e,f) raised the question to what extent this network- to-neuron impact is regulated by a neuron's in-degree, i.e., the number of input connections to a neuron. To address this question, we took advantage of the distri- bution of in-degrees provided by a model network with sparse and random connectivity. We found that near criticality (λ ≈ 1), a neuron's CV tended to increase with increasing in-degree (fig.2a). Im- portantly, however, this correlation between a neuron's activity statistics (CV) and its connectivity (in-degree) changed drastically when tuning the network state away from criticality. In the subcritical regime only weak cor- relation was found and in the supercritical regime the relation reversed, i.e., a neuron's CV tended to decrease with increasing in-degree (fig.2b). In general, when tun- ing the network state through the critical regime, the relation between a neuron's CV and its in-degree trans- forms from small correlation in the subcritical regime (λ < 1), to maximized correlation at the critical point (λ ≈ 1), and to anti-correlation in the supercritical regime (λ > 1), (fig.2c). The relationship that solely near criticality is the ir- regular spiking maximally reflective of the neuron's in- degree, suggests a novel measure to test the criticality hypothesis in electrophysiological experiments. It is of practical importance that in a network dominated by ex- citatory neurons, a neuron's firing rate scales with its in-degree and that this relation is independent of the network state. Thus, spike train recordings from a pop- ulation of neurons can be informative about the network state: a maximum correlation between a neuron's CV and its firing rate is indicative of a critical network state, whereas weaker correlation or anti-correlation is indica- tive of the subcritical or supercritical network state, re- spectively (fig.2d-f). In comparison, the relation between a neuron's "pop- ulation coupling" (see Methods) and its in-degree is less decisive about the network state, as population coupling increases with a neuron's in degree for all three network states (fig.2g, h). However, this relation is also dominant 4 Fig. 3: The network state determines the change in response variability after stimulus onset. (a) Average Fano factor (solid lines) from repeated step increases in the external input for three different network states: subcritical (λ = 0.95, blue), critical (λ = 1.02, red), and supercritical (λ = 1.07, green). Spike trains were simulated for a network with N = 5000 neurons and 3% connectivity. External input was applied to 10% of the neurons and external input strength switched from η = 1/N to η = 5/N as indicated. The Fano factor is the ratio of the variance and mean in the number of spikes within a given time window and repeated trials. The sliding window was 200 time steps and the sliding increment was 20 time steps. We calculated the average Fano factor from 60 randomly sampled model neurons and 2000 repeated trials. (b) The mean and variance of the spike rate when the average external input strength switched from 1/N (first half) to η = 5/N (second half) for three different network states. All network parameters as in (a). for the critical network state (fig.2i). Consistent with these model results, recent experimental data from in vivo recordings showed increased population coupling with increasing synaptic inputs [11]. However, unless the network state is manipulated and a maximum in the correlation between the population coupling and the in-degree is determined, that data set is not informative about the network state. In addition to the irregular spiking evaluated above, a decline of response variability at stimulus onset is a prominent cortical phenomenon [3]. The important ques- tion that how an external input suppresses the variability of ongoing activity has remained intriguing [12, 13]. To address this question in the context of the network state, we simulated spiking activity of the model network near criticality (λ ≈ 1) in response to repeated increases in the external input. We quantified the across-trial firing rate variability in terms of the Fano factor, which is a mea- sure of the across-trial variability in the number of spikes in relation to the trial-averaged mean number of spikes during a given window of time (see Methods). The Fano factor has a value of approximately 1 for repeated Pois- son spike trains. Recordings from cortical neurons show a Fano factor above 1 for ongoing activity and a significant drop during external stimulation [3]. Our simulations of a model network in the critical state reproduced this drop in Fano factor from a high value during ongoing activity to a lower value (still above 1) during external stimulation (fig.3a). In contrast, in the subcritical and the supercrit- ical network state, the Fano factor remained unchanged when switching from ongoing to evoked activity. This re- sult demonstrates that the experimentally observed drop in Fano factor could naturally emerge as a characteristic feature of the critical network state. The result is quali- tatively robust with respect to the details of the external inputs (fig.S4). Since the Fano factor is defined as a ratio of the vari- ance divided by the mean, the declining Fano factor at criticality could result trivially from an increased mean spike count accompanied by a weak dependence of the variance on the network state during external stimula- tion. Evaluating the changes in the spike count and the variance separately ruled out this possibility (fig.3b). The "mean spike count" and the variance increased with external stimulation for all three network states. How- ever, only near criticality, when the network is maximally sensitive to external stimuli, did the mean spike count in- crease more drastically, thus yielding a reduced Fano fac- tor. In conclusion, our model simulation suggests that the experimentally observed drop in Fano factor after stimulus onset [3] can be the result of the cortical circuit operating at criticality. III. Discussion A wealth of evidence indicates that neural networks of cerebral cortex operate at criticality [4, 14, 15]. Here, we showed that (i) irregular spiking [2, 6, 16], (ii) its relation to the neuron's in-degree, and (iii) the decline of response variability at stimulus onset [3], are all emergent properties of a recurrent network operating at criticality. A number of separate dynamical, biophysical, and structural mechanisms have been proposed to generate the observed irregular spiking [12, 17 -- 21]. The significance of our work resides in part in establishing criticality as one unifying principle for both the collective behavior of cortical circuits and the statistics of single- neuron spiking. Indeed, experimental evidence for the predicted coexistence of irregular spiking and criticality has recently been provided in recordings of ongoing cortical activity in vivo [6]. While the coexistence of irregular spiking and power-law avalanche size distri- butions has been demonstrated in more complex model networks[22 -- 26], our work extends qualitatively beyond these important earlier studies in three fundamental dimensions. First, the choice of a network of excitatory probabilistic integrate-and-fire model neurons allows the precise analytic evaluation of the network state via a single control parameter λ, i.e., the maximum eigenvalue of the transfer probability matrix. This approach avoids the need to rely on the cumbersome and less precise avalanche analysis to evaluate the network state. Second, our discovery that the relation between the irregularity of spiking and the neuron's in-degree/firing rate is maximized at criticality provides an important new measure of criticality. Furthermore, this relation establishes a valuable conceptual link between criticality and the important field of network theory, where a node's in-degree is a basic system parameter. Third, as the observed decline in response variability is regarded as an essential mechanism to enhance response fidelity to stimuli [3], our discovery of its relation to network criticality offers a starting point toward unraveling the possible roles of critical dynamics in neural coding. In conclusion, it will be interesting to see to what extent the presented findings will generate a paradigm shift in the study of criticality of neural systems: our results build a much-needed bridge between critical dynamics and neural coding, and provide novel and robust measures to test the criticality hypothesis itself. IV. Methods We simulated a model network consisting of excitatory bi- nary probabilistic model neurons with sparse connectivity and external inputs. Network size ranged from 5000 to 20000 model neurons. The strength of the connection from neuron j to neuron i is quantified in terms of the transition probability Pij , which is the probability that a spike in neuron j causes a spike in neuron i in the next simulation time step. For a network of N neurons and an average connectivity K, each neuron is connected to N − 1 other neurons with probability K/N . For each (on average) K(N − 1) connections among neurons a Pij is assigned by drawing a random number from uniform distribution in the interval [0 2 K ]. With sufficiently 5 large this yields a network with a normally distributed con- nectivity with average K and a transition matrix Pij with maximum (absolute value) eigenvalue of 1. In order to devi- ate the network from the critical point we can simply multiply Pij by a factor smaller(greater) than one. The binary state Xi(t) of neuron i denotes whether the model neuron spikes (Xi(t) = 1 or does not spike (Xi(t) = 0 at time t. At each time step, the state of all neurons were updated synchronously according to the following update rule: Xi(t + 1) = Θ(cid:20)(cid:16)1 − ηi(t)(cid:17)Xj Pij Xj(t) + ηi(t) − ξi(t)(cid:21) (1) where ηi(t) of neuron i quantifies the probability of that neuron spiking only due to external input, ξ(t) is a random number in [01] drawn from a uniform distribution, and Θ is the step function. In addition to the update rule, a refractory period of two time-steps was implemented. The external input ηi(t) was chosen to be smaller than the transition probabil- ity Pij , which itself was small for large networks, Pij ∼ 1/N . Because of the weak external inputs, we employed the approx- imation 1 − η ≈ 1 in the above update rule. The maximum eigenvalue λ of the transition probability matrix Pij describes the network state at the thermodynamic limit: λ < 1 denotes subcritical regime, λ ≈ 1 denotes the near critical regime and λ > 1 denotes the supercritical regime. However, for finite- sized networks we evaluate the exact critical point based on the peak of average population coupling (see below). The Pij values were drawn from a uniform distribution and then scaled by constant to reach the desired maximum eigenvalue λ. The constant external input ηi(t) was either constant. The external input ηi(t) was either constant or was mod- eled as a binary Poisson process followed by smoothing with a Gaussian filter with a width of 20 time steps and multiplied by an amplitude factor η0 between 0 and 1. The synchronous input (fig.S1) was simulated with replicating a single binary Poisson process smoothed with a Gaussian filter with a width of 100 time-steps. In order to apply some variability among the stimuli received by different neurons each smoothed Pois- son process was multiplied by a factor of η0 + 0.2ǫ where ǫ was drawn from a normal distribution. We analyzed the simulated spike trains with respect to five complementary statistical measures: neuronal avalanches, coefficient of variation, population coupling, change in mean response, and Fano factor. Following commonly-used proce- dures [27], a neuronal avalanche was defined as an episode of continuous (each time step) network spiking, framed by time steps of no spikes. Avalanche size was taken as the number of spiking neurons. The single-neuron spike train variability was quantified using the coefficient of variation CV ≡ σisi/µisi, defined as the ratio of the standard deviation σisi and the mean µisi of the inter-spike-interval (ISI) distribution for a given neuron. We managed simulation times to be sufficiently long to ensure stable CV values. The coordination within the network was quantified using the population coupling , which is defined as the zero-lag cross-correlation between the spike train Xi(t) of neuron i and the remaining network activity Ni(t) = Pj6=i Xj (t) from all other spike trains, i.e., ci = D(cid:0)Xi(t)−hXi(t)i(cid:1)(cid:0)Ni(t)−hNi(t)i(cid:1)E , where the angled brack- ets indicate a time average [11]. Averaging the population coupling over many neurons within a large network yields the average population coupling , which represents a measure of the overall level of coordination within the network. The σX σN change in mean response was computed as the difference in the mean spike counts for external inputs of η = 1/(5N ) and η = 2/N . When quantifying the spike train variability in the context of repeated experimental situations, it is convenient to use the Fano factor, which, for a given time window and repeated trials, is defined as the ratio of the variance and mean in the number of spikes. We chose a window of 200 time-steps (sliding in increments of 20 time steps) and repeated trials 2000 times. We calculated the average Fano factor from 60 randomly sampled model neurons. Acknowledgment We thank Anders Carlsson, John Clark, and Woodrow Shew for comments on previous ver- sions of the manuscript. This research was supported by a Whitehall Foundation grant #20121221 (R.W.) and a NSF CRCNS grant #1308159 (R.W.). 6 [1] H. Sompolinsky, "Computational neuroscience: beyond the local circuit," Current opinion in neurobiology, vol. 25, pp. xiii -- xviii, 2014. [2] W. R. Softky and C. Koch, "The highly irregular firing of cortical cells is inconsistent with temporal integration of random epsps," The Journal of Neuroscience, vol. 13, no. 1, pp. 334 -- 350, 1993. [3] M. M. Churchland, M. Y. Byron, J. P. Cunningham, L. P. Sugrue, M. R. Cohen, G. S. Corrado, W. T. Newsome, A. M. Clark, P. Hosseini, B. B. Scott, et al., "Stimu- lus onset quenches neural variability: a widespread cor- tical phenomenon," Nature neuroscience, vol. 13, no. 3, pp. 369 -- 378, 2010. [4] H. G. Schuster, D. Plenz, and E. Niebur, Criticality in neural systems. John Wiley & Sons, 2014. [5] P. Bak, C. Tang, and K. Wiesenfeld, "Self-organized crit- icality: An explanation of the 1/f noise," Physical review letters, vol. 59, no. 4, p. 381, 1987. [6] T. Bellay, A. Klaus, S. Seshadri, and D. Plenz, "Irregular spiking of pyramidal neurons organizes as scale-invariant neuronal avalanches in the awake state," eLife, vol. 4, p. e07224, 2015. [7] S. Denève and C. K. Machens, "Efficient codes and bal- anced networks," Nature neuroscience, vol. 19, no. 3, pp. 375 -- 382, 2016. [8] W. Gerstner, W. M. Kistler, R. Naud, and L. Panin- ski, Neuronal dynamics: From single neurons to networks and models of cognition. Cambridge University Press, 2014. [9] D. B. Larremore, W. L. Shew, and J. G. Restrepo, "Predicting criticality and dynamic range in complex networks: effects of topology," Physical review letters, vol. 106, no. 5, p. 058101, 2011. [10] D. B. Larremore, M. Y. Carpenter, E. Ott, and J. G. Re- strepo, "Statistical properties of avalanches in networks," Physical Review E, vol. 85, no. 6, p. 066131, 2012. [11] M. Okun, N. A. Steinmetz, L. Cossell, M. F. Iacaruso, H. Ko, P. Barthó, T. Moore, S. B. Hofer, T. D. Mrsic- Flogel, M. Carandini, et al., "Diverse coupling of neu- rons to populations in sensory cortex," Nature, vol. 521, no. 7553, pp. 511 -- 515, 2015. [12] D. Sussillo and L. F. Abbott, "Generating coherent pat- terns of activity from chaotic neural networks," Neuron, vol. 63, no. 4, pp. 544 -- 557, 2009. [13] K. Rajan, L. Abbott, and H. Sompolinsky, "Stimulus- dependent suppression of chaos in recurrent neural net- works," Physical Review E, vol. 82, no. 1, p. 011903, 2010. [14] O. Arviv, A. Goldstein, and O. Shriki, "Near-critical dy- namics in stimulus-evoked activity of the human brain and its relation to spontaneous resting-state activity," The Journal of Neuroscience, vol. 35, no. 41, pp. 13927 -- 13942, 2015. [15] W. L. Shew, W. P. Clawson, J. Pobst, Y. Karimipanah, N. C. Wright, and R. Wessel, "Adaptation to sensory in- put tunes visual cortex to criticality," Nature Physics, vol. 11, no. 8, pp. 659 -- 663, 2015. [16] M. N. Shadlen and W. T. Newsome, "The variable dis- charge of cortical neurons: implications for connectivity, computation, and information coding," The Journal of neuroscience, vol. 18, no. 10, pp. 3870 -- 3896, 1998. [17] C. van Vreeswijk and H. Sompolinsky, "Chaos in neuronal networks with balanced excitatory and inhibitory activ- ity," Science, vol. 274, no. 5293, pp. 1724 -- 1726, 1996. [18] C. F. Stevens and A. M. Zador, "Input synchrony and the irregular firing of cortical neurons," Nature neuroscience, vol. 1, no. 3, pp. 210 -- 217, 1998. [19] A. Kumar, S. Rotter, and A. Aertsen, "Conditions for propagating synchronous spiking and asynchronous fir- ing rates in a cortical network model," The Journal of neuroscience, vol. 28, no. 20, pp. 5268 -- 5280, 2008. [20] A. Litwin-Kumar and B. Doiron, "Slow dynamics and high variability in balanced cortical networks with clus- tered connections," Nature neuroscience, vol. 15, no. 11, pp. 1498 -- 1505, 2012. [21] S. Ostojic, "Two types of asynchronous activity in net- works of excitatory and inhibitory spiking neurons," Na- ture neuroscience, vol. 17, no. 4, pp. 594 -- 600, 2014. [22] D.-M. Chen, S. Wu, A. Guo, and Z. Yang, "Self-organized criticality in a cellular automaton model of pulse-coupled integrate-and-fire neurons," Journal of physics A: math- ematical and general, vol. 28, no. 18, p. 5177, 1995. [23] C. W. Eurich, J. M. Herrmann, and U. A. Ernst, "Finite- size effects of avalanche dynamics," Physical review E, vol. 66, no. 6, p. 066137, 2002. [24] M. Benayoun, J. D. Cowan, W. van Drongelen, and E. Wallace, "Avalanches in a stochastic model of spiking neurons," PLoS Comput Biol, vol. 6, no. 7, p. e1000846, 2010. [25] D. Millman, S. Mihalas, A. Kirkwood, and E. Niebur, "Self-organized criticality occurs in non-conservative neu- ronal networks during/up/'states," Nature physics, vol. 6, no. 10, pp. 801 -- 805, 2010. [26] N. Stepp, D. Plenz, and N. Srinivasa, "Synaptic plastic- ity enables adaptive self-tuning critical networks," PLoS Comput Biol, vol. 11, no. 1, p. e1004043, 2015. [27] J. M. Beggs and D. Plenz, "Neuronal avalanches in neo- cortical circuits," The Journal of neuroscience, vol. 23, no. 35, pp. 11167 -- 11177, 2003. [28] J. M. Beggs and N. Timme, "Being critical of criticality in the brain," Front Physiol, vol. 3, p. 163, 2012. Supplementary Figures 7 Fig. S1: Irregular spiking at criticality in a recurrent network with synchronous external inputs. (a) The temporal structure and strength of the external input η(t) to 10% of the neurons in a recurrent model network of 5000 neurons and 1% connectivity. The external input η(t) was generated from Poisson pulses of rate 10/N , smoothed by a Gaussian filter of width 100 time-steps and amplitude of 0.2(1 + ǫ), where ǫ is drawn from a normal distribution (see Methods). This synchronous input was added to a background constant external input of 1/(10N ). (b) Inter-spike-interval CV distributions of simulated spike trains for the subcritical (λ = 0.95, blue), critical (λ = 1.09, red), and supercritical (λ = 1.09, green) network state. At the critical regime the spike trains show highest irregularity, which is indicated by the peak of the CV distribution located near 1.3. (c, d) The inter-spike-interval CVs from simulated spike trains versus the neuron's in degree (c) and its normalized rate (d) for the three network states. (e, f ) The population coupling from simulated spike trains versus the neuron's in degree (e) and its normalized rate (f) for the three network states. (g) The population coupling vs a neuron's CV for the three network states. (h) The Spearman correlation coefficients between CV and in-degree (rate), population coupling and in-degree (rate), population coupling and CV are all maximized near criticality. 8 Fig. S2: Irregular spiking at criticality in a recurrent network with asynchronous external inputs. (a) The temporal structure and strength of the external input ǫ to 10% of the neurons in a recurrent model network of 5000 neurons and 1% connectivity. The external input ǫ was generated by independent Poisson pulses of rate 5/N , smoothed by a Gaussian filter of width 20 time-steps and amplitude η0 = 0.5 (see Methods). (b) Inter-spike-interval CV distributions of simulated spike trains for the subcritical (λ = 0.95, blue), critical (λ = 1.02, red), and supercritical (λ = 1.09, green) network state. (c, d) The inter-spike-interval CVs from simulated spike trains versus the neuron's in degree (c) and its normalized rate (d) for the three network states. (e, f ) The population coupling from simulated spike trains versus the neuron's in degree (e) and its normalized rate (f) for the three network states. (g) The population coupling vs a neuron's CV for the three network states. (h) The Spearman correlation coefficients between CV and in-degree (rate), population coupling and in-degree (rate), population coupling and CV are all maximized near criticality. 9 synchronized stimuli synchronized stimuli synchronized stimuli synchronized stimuli a c e b d f random stimuli random stimuli Fig. S3: The average CV and average population coupling are maximized near network criticality for external inputs of different spatiotemporal structure. (a, b) Average CV (a) and average population coupling (b) vs the control parameter ?? for synchronous external inputs (see Fig. S1a, but with different stimulation amplitudes (see fig.S1a, but with stimulation amplitude η0 = 0.1, 0.2, 0.5; see Methods) for three different network sizes. (c, d) Average CV (c) and average population coupling (d) vs the control parameter λ for synchronous external inputs (see fig.S1a) for a network size of N = 5000 and for three different stimulus amplitudes. (e, f ) Average CV (e) and average population coupling (f) vs the control parameter λ for asynchronous external inputs (see fig.S2a) for three different network sizes (see legend in (a)). 10 Fig. S4: The change in response variability for external inputs of different temporal structure. (a) Average external input for the case of synchronous external inputs applied to 10% of the neurons as described in fig.S1a. (b) The average Fano factor computed for 60 neurons over 2000 trials of different stimuli, but of the same type. In the presence of temporally structured stimuli, the Fano factor can increase at sub-criticality, due to the across-trial variability in the stimuli. (c) The average Fano factor computed for the same neurons over 2000 trials of the exactly the same stimulus (as is shown in (a). (d-f ) Same as (a-b) but for asynchronous external inputs as described in fig.S2a.
1809.08045
3
1809
2019-08-24T12:58:19
Stochasticity from function -- why the Bayesian brain may need no noise
[ "q-bio.NC", "cond-mat.dis-nn", "cs.NE", "physics.bio-ph", "stat.ML" ]
An increasing body of evidence suggests that the trial-to-trial variability of spiking activity in the brain is not mere noise, but rather the reflection of a sampling-based encoding scheme for probabilistic computing. Since the precise statistical properties of neural activity are important in this context, many models assume an ad-hoc source of well-behaved, explicit noise, either on the input or on the output side of single neuron dynamics, most often assuming an independent Poisson process in either case. However, these assumptions are somewhat problematic: neighboring neurons tend to share receptive fields, rendering both their input and their output correlated; at the same time, neurons are known to behave largely deterministically, as a function of their membrane potential and conductance. We suggest that spiking neural networks may, in fact, have no need for noise to perform sampling-based Bayesian inference. We study analytically the effect of auto- and cross-correlations in functionally Bayesian spiking networks and demonstrate how their effect translates to synaptic interaction strengths, rendering them controllable through synaptic plasticity. This allows even small ensembles of interconnected deterministic spiking networks to simultaneously and co-dependently shape their output activity through learning, enabling them to perform complex Bayesian computation without any need for noise, which we demonstrate in silico, both in classical simulation and in neuromorphic emulation. These results close a gap between the abstract models and the biology of functionally Bayesian spiking networks, effectively reducing the architectural constraints imposed on physical neural substrates required to perform probabilistic computing, be they biological or artificial.
q-bio.NC
q-bio
Stochasticityfromfunction -- whytheBayesianbrainmayneednonoise DominikDolda,b,1,2,IljaBytschoka,1,AkosF.Kungla,b,AndreasBaumbacha, OliverBreitwiesera,WalterSennb,JohannesSchemmela,KarlheinzMeiera,MihaiA.Petrovicia,b,1,2 aKirchhoff-InstituteforPhysics,HeidelbergUniversity. bDepartmentofPhysiology,UniversityofBern. 1Authorswithequalcontributions. 2Correspondingauthors:[email protected],[email protected]. August27,2019 Abstract Anincreasingbodyofevidencesuggeststhatthetrial-to-trialvariabilityofspikingactivityinthebrainisnotmerenoise,but ratherthereflectionofasampling-basedencodingschemeforprobabilisticcomputing.Sincetheprecisestatisticalproperties ofneuralactivityareimportantinthiscontext,manymodelsassumeanad-hocsourceofwell-behaved,explicitnoise,either ontheinputorontheoutputsideofsingleneurondynamics,mostoftenassuminganindependentPoissonprocessineither case.However,theseassumptionsaresomewhatproblematic:neighboringneuronstendtosharereceptivefields,rendering boththeirinputandtheiroutputcorrelated;atthesametime,neuronsareknowntobehavelargelydeterministically,asa functionoftheirmembranepotentialandconductance.Wesuggestthatspikingneuralnetworksmayhavenoneedfornoise toperformsampling-basedBayesianinference.Westudyanalyticallytheeffectofauto-andcross-correlationsinfunctional Bayesianspikingnetworksanddemonstratehowtheireffecttranslatestosynapticinteractionstrengths,renderingthem controllablethroughsynapticplasticity.Thisallowsevensmallensemblesofinterconnecteddeterministicspikingnetworksto simultaneouslyandco-dependentlyshapetheiroutputactivitythroughlearning,enablingthemtoperformcomplexBayesian computationwithoutanyneedfornoise,whichwedemonstrateinsilico,bothinclassicalsimulationandinneuromorphic emulation.TheseresultscloseagapbetweentheabstractmodelsandthebiologyoffunctionallyBayesianspikingnetworks, effectively reducing the architectural constraints imposed on physical neural substrates required to perform probabilistic computing,betheybiologicalorartificial. 1 9 1 0 2 g u A 4 2 ] . C N o i b - q [ 3 v 5 4 0 8 0 . 9 0 8 1 : v i X r a 1 Introduction An ubiquitous feature of in-vivo neural responses is their stochasticnature[1 -- 6].Theclearpresenceofthisvariabil- ityhasspawnedmanyfunctionalinterpretations,withthe Bayesian-brainhypothesisarguablybeingthemostnotable example[7 -- 12].Underthisassumption,theactivityofaneu- ral network is interpreted as representing an underlying (prior)probabilitydistribution,withsensorydataproviding theevidenceneededtoconstrainthisdistributiontoa(pos- terior)shapethatmostaccuratelyrepresentsthepossible statesoftheenvironmentgiventhelimitedavailableknowl- edgeaboutit. Neural network models have evolved to reproduce this kindofneuronalresponsevariabilitybyintroducingnoise- generatingmechanisms,betheyextrinsic,suchasPoisson input [13 -- 16] or fluctuating currents [17 -- 22], or intrinsic, suchasstochasticfiring[23 -- 28]ormembranefluctuations [29,30].However, while representing, to some degree, reason- ableapproximations,noneofthecommonlyusedsourcesof stochasticityisfullycompatiblewithbiologicalconstraints. Contrarytotheindependentwhitenoiseassumption,neu- ronal inputsare bothauto- and cross-correlatedto asig- nificant degree [31 -- 37], with obvious consequences for a network's output statistics [38]. At the same time, the as- sumptionofintrinsicneuronalstochasticityisatoddswith experimentalevidenceofneuronsbeinglargelydeterminis- ticunits[39 -- 41].Althoughsynaptictransmissionsfromin- dividualreleasesitesarestochastic,averagedacrossmul- tiplesites,contactsandconnections,theylargelyaverage out[42].Therefore,itremainsaninterestingquestionhow corticalnetworksthatusestochasticactivityasameansto performprobabilisticinferencecanrealisticallyattainsuch apparentrandomnessinthefirstplace. We address this question within the normative frame- workofsampling-basedBayesiancomputation[29,43 -- 47], in which the spiking activity of neurons is interpreted as Markov Chain Monte Carlo sampling from an underlying distributionoverahigh-dimensionalbinarystatespace.In contrasttootherworkondeterministicchaosinfunctional spiking networks, done mostly in the context of reservoir computing (e.g., [48, 49]), we provide a stringent connec- tiontothespike-basedrepresentationandcomputationof probabilities,aswellasthesynapticplasticityrequiredfor learning the above. We demonstrate how an ensemble of dynamicallyfullydeterministic,butfunctionallyprobabilis- ticnetworks,canlearnaconnectivitypatternthatenables probabilistic computation with a degree of precision that matchestheoneattainablewithidealized,perfectlystochas- ticcomponents.Thekeyelementofthisconstructionisself- consistency,inthatallinputactivityseenbyaneuronisthe resultofoutputactivityofotherneuronsthatfulfillafunc- tionalroleintheirrespectivesubnetworks.Thepresentwork supportsprobabilisticcomputationinlightofexperimental 2 (cid:88) gsyn k,x (t) = synapses j spikes s Erev (cid:88) x∈{e,i} Cm duk dt = gl (El − uk) + gsyn k,x (Erev wkj θ(t − ts) exp (− x − uk) , t − ts τ syn ) , uk(ts) ≥ ϑ ⇒ uk(t ∈ (ts, ts + τref ]) =  , evidencefrombiologyandsuggestsaresource-efficientim- plementationofstochasticcomputingbycompletelyremov- ingtheneedforanyformofexplicitnoise. 2 Methods 2.1 Neuronmodelandsimulationdetails We consider deterministic Leaky Integrate-and-Fire (LIF) neurons with conductance-based synapses and dynamics describedby (1) (cid:88) (2) (3) withmembranecapacitance Cm,leakconductance gl,leak potential El, excitatory and inhibitory reversal potentials e/i andconductances gsyn k,e/i,synapticstrength wkj,synap- tictimeconstant τ synand θ(t)theHeavisidestepfunction. For gsyn k,e/i,thefirstsumcoversallsynapticconnectionspro- jecting to neuron k. A neuron spikes at time ts when its membranepotentialcrossesthethreshold ϑ,afterwhichit becomes refractory. During the refractory period τref, the membranepotentialisclampedtotheresetpotential .We havechosentheabovemodelbecauseitprovidesacompu- tationallytractableabstractionofneurosynapticdynamics [41],butourgeneralconclusionsarenotrestrictedtothese specificdynamics. We further use the short-term plasticity mechanism de- scribed in [50] to modulate synaptic interaction strengths with an adaptive factor USE × R(t), where the time- dependenceisgivenby1 (4) where δ(t)istheDiracdeltafunction, ts denotesthetime ofapresynapticspike,whichdepletesthereservoir Rbya fraction USE,and τrecisthetimescaleonwhichthereser- voir Rrecovers.Thisenablesabettercontrolovertheinter- neuroninteraction,aswellasoverthemixingpropertiesof ournetworks[47]. Backgroundinput,suchasspikesfromaPoissonsource, entersEq.(1)assynapticinput,butwithoutshort-termplas- ticity(asin[44])tofacilitatethemathematicalanalysis(see Supportinginformationformoredetails). All simulations were performed with the network spec- ification language PyNN 0.8 [51] and the spiking neural networksimulatorNEST2.4.2[52]. 1In[50]thepostsynapticresponseonlyscaleswith R(t),whereashere wescaleitwith USE × R(t). 1 − R τrec − USERδ(t − ts) , USE , R ∈ [0, 1] , dR dt = ptarget z τ syn e− τref τ syn τref αWkjCm τ syn − 1 (cid:17)−1 (cid:17)(cid:105) , (cid:17) 1 − τ syn (cid:88) − τeff = p(z) ∝ exp(cid:0) 1 (cid:16) (cid:104) 2.2 Samplingframework As a model of probabilistic inference in networks of spik- ingneurons,weadopttheframeworkintroducedin[44,46]. There,theneuronaloutputbecomesstochasticduetoahigh- frequencybombardmentofexcitatoryandinhibitoryPoisson stimuli(Fig.1A),elevatingneuronsintoahigh-conductance state(HCS)[53,54],wheretheyattainahighreactionspeed duetoareducedeffectivemembranetimeconstant.Under theseconditions,aneuron'sresponse(oractivation)function becomesapproximatelylogisticandcanberepresentedas ϕ(µ) = (1 + exp (−(µ − u0)/α))−1 with inverse slope α andinflectionpoint u0.Togetherwiththemeanfreemem- brane potential µ and the mean effective membrane time constant τeff (seeEqs.(17b)and(20c)),thescalingparam- eters α and u0 are used to translate the weight matrix W and bias vector b of a target Boltzmann distribution 2 zT W z + zT b(cid:1)withbinaryrandom variables z ∈ {0, 1}n tosynapticweightsandleakpoten- tialsinasamplingspikingnetwork(SSN): (cid:16) (cid:17)(cid:16) (5) (cid:16) Erev kj − µ (6) e− τeff − 1 (cid:104)gsyn x (cid:105) Erev where wkjisthesynapticweightfromneuron jtoneuron k, gl Elavectorcontainingtheleakpotentialsofallneurons, b i },depend- thecorrespondingbiasvector, Erev ingonthenatureoftherespectivesynapse,and τm = Cm (seeEq.(68)toEq.(73)foraderivation).Thistranslation effectivelyenablessamplingfrom ptarget ,wherearefractory neuron is considered to represent the state zk = 1 (see Fig.1B,C). 2.3 Measuresofnetworkperformance Toassesshowwellasamplingspikingnetwork(SSN)sam- plesfromitstargetdistribution,weusetheKullback-Leibler divergence[55] (7) whichisameasureforthesimilaritybetweenthesampled distribution pnet andthetargetdistribution ptarget.Forin- ferencetasks,wedeterminethenetwork'sclassificationrate onasubsetoftheuseddatasetwhichwasputasidedur- ing training. Furthermore, generative properties of SSNs areinvestigatedeitherbylettingthenetworkcompletepar- tiallyoccludedexamplesfromthedatasetorbylettingit generatenewexamples. (cid:0)pnet (cid:107) ptarget(cid:1) = (cid:18) pnet (αb + u0) − kj ∈ {Erev (cid:88) pnet z ln z ptarget z , Erev e (cid:19) , El = τm τeff τref , x τeff x∈{e,i} DKL z wkj = gl z , (cid:105) (cid:104) zi=1,zj =1 (cid:3) , ∆W ij = η ptarget zi=1,zj =1 − pnet zi=1 − pnet ∆bi = η(cid:2)ptarget Figure 1: Sampling spiking networks (SSNs) with and without explicit noise. (A) Schematic of a sampling spiking network, where each neuron (circles) encodes a binary random variable zi ∈ {0, 1}. In the original model, neurons were rendered effectively stochastic by adding external Poissonsourcesofhigh-frequencybalancednoise(redboxes).(B)Aneuron representsthestate zk = 1whenrefractoryand zk = 0otherwise.(C) ThedynamicsofneuronsinanSSNcanbedescribedassampling(red bars)fromatargetdistribution(bluebars).(D)InsteadofusingPoisson processesasasourceofexplicitnoise,wereplacethePoissoninputwith spikes coming from other networks performing spike-based probabilistic inference by creating a sparse, asymmetric connectivity matrix between severalSSNs.Forinstance,theredneuronreceivesnotonlyinformation- carryingspikesfromitshomenetwork(blacklines),butalsospikesfrom theothertwoSSNsasbackground(redarrows),andinturnprojectsback towardsthesenetworks.Othersuchbackgroundconnectionsareindicated inlightgray. 2.4 Learningalgorithm Networks were trained with a Hebbian wake-sleep algo- rithm (8) (9) whichminimizesthe DKL (pnet (cid:107) ptarget)[56]. η isalearn- ingrate(seeSupportinginformationforusedhyperparam- eters).Forhigh-dimensionaldatasets(e.g.handwrittenlet- tersanddigits),Boltzmannmachinesweretrainedwiththe CASTalgorithm[57],avariantofwake-sleepwithatemper- ing scheme, and then translated to SSN parameters with Eqs. (5) and (6) instead of training the SSNs directly to reducesimulationtime. 2.5 Experimentsandcalculations Detailstoallexperimentsaswellasadditionalfiguresand captionstovideoscanbefoundintheSupportinginforma- tion.Detailedcalculationsarepresentedattheendofthe maintext. 3 Results We approach the problem of externally-induced stochas- ticity incrementally. Throughout the manuscript, we discern between background input, which isprovidedbyotherfunctionalnetworks,andexplicitnoise, for which we use the conventional assumption of Poisson the remainder of zi=1 3 t [a.u.]u2u1u0101100zB000001010011100101110111z102101p(z)CADz0z1z2 spiketrains.Westartbyanalyzingtheeffectofcorrelated backgroundontheperformanceofSSNs.Wethendemon- strate how the effects of both auto- and cross-correlated background can be mitigated by Hebbian plasticity. This ultimately enables us to train a fully deterministic net- workofnetworkstoperformdifferentinferencetaskswith- outrequiringanyformofexplicitnoise.Thisisfirstshown for larger ensembles of small networks, each of which re- ceivesitsowntargetdistribution,whichallowsastraightfor- wardquantitativeassessmentoftheirsamplingperformance DKL (pnet (cid:107) ptarget).Westudythebehaviorofsuchensem- blesbothincomputersimulationsandonmixed-signalneu- romorphichardware.Finally,wedemonstratethecapability ofourapproachfortrulyfunctional,larger-scalenetworks, trainedonhigher-dimensionalvisualdata. 3.1 Backgroundautocorrelations Unlike ideal Poisson sources, single spiking neurons pro- duceautocorrelatedspiketrains,withtheshapeoftheau- tocorrelationfunction(ACF)dependingontheirrefractory time τref and mean spike frequency ¯r = p(z = 1)τref−1. For higher output rates, spike trains become increasingly dominated by bursts, i.e., sequences of equidistant spikes with an interspike interval (ISI) of ISI ≈ τref. These fixed structures also remain in a population, since the popula- tionautocorrelationisequaltotheaveragedACFsofthe individualspiketrains. We investigated the effect of such autocorrelations on the output statistics of SSNs by replacing the Poisson input in the ideal model with spikes coming from other SSNs. As opposed to Poisson noise, the autocorrelation oftheSSN-generated (excitatoryorinhibitory)background Sx, x ∈ {e, i}(Fig.2B) isnon-singularandinfluencesthefreemembranepotential C(Sx, Sx, ∆) = (cid:104)Sx(t)Sx(t+∆)(cid:105)−(cid:104)Sx(cid:105)2 (FMP)distribution(Fig.2C)andtherebyactivationfunction (Fig. 2D) of individual sampling neurons. With increasing firing rates (controlled by the bias of the neurons in the backgroundSSNs),thenumberofsignificantpeaksinthe ACFincreasesaswell(seeEq.(54)): (10) where ¯pistheprobabilityforabursttostart.Thisregularity inthebackgroundinputmanifestsitselfinareducedwidth σ(cid:48)oftheFMPdistribution(seeEq.(30)) (cid:112) (11) withascalingfactor √βthatdependsontheACF,whichin turntranslatestoasteeperactivationfunction(seeEqs.(36) and(37)) (12) ek ln ¯pδ(cid:0)[n − k]τref ) ∼ N (µ(cid:48) = µ, σ(cid:48) = C(Sx, Sx, nτref ) ≈ (cid:90) ∞ ∞(cid:88) (cid:1) , f (ufree Var(Sx) βσ) k=1 i , f (u)du ≈ ϕ(µ) p(zi = 1) ≈ ϑ (cid:12)(cid:12)(cid:12)(cid:12)u(cid:48) 0=u0,α(cid:48)=√βα withinflectionpoint u(cid:48)0andinverseslope α(cid:48).Thus,autocor- relationsinthebackgroundinputleadtoareducedwidth oftheFMPdistributionandhencetoasteeperactivation functioncomparedtotheoneobtainedusinguncorrelated Poisson input. For a better intuition, we used an approx- imation of the activation function of LIF neurons, but the argumentalsoholdsfortheexactexpressionderivedin[44], asverifiedbysimulations(Fig.2D). Apartfromtheaboveeffect,thebackgroundautocorrela- tionsdonotaffectneuronpropertiesthatdependlinearly on the synaptic noise input, such as the mean FMP and the inflection point of the activation function (equivalent tozerobias).Therefore,theeffectofthebackgroundauto- correlations can be functionally reversed by rescaling the functional(fromotherneuronsintheprincipalSSN)affer- entsynapticweightsbyafactorequaltotheratiobetween thenewandtheoriginalslope α(cid:48)/α(Eqs.(5)and(6)),as showninFig.2E. 3.2 Backgroundcross-correlations In addition to being autocorrelated, background input to pairs of neurons can be cross-correlated as well, due to either shared inputs or synaptic connections between the neuronsthatgeneratesaidbackground.Thesebackground cross-correlations can manifest themselves in a modified cross-correlation between the outputs of neurons, thereby distortingthedistributionsampledbyanSSN. However,dependingonthenumberandnatureofpresy- naptic background sources, background cross-correlations maycancelouttoasignificantdegree.Thecorrelationco- efficient(CC)oftheFMPsoftwoneuronsfedbycorrelated noiseamountsto(seeEq.(59)) (cid:1) (13) (cid:0)Erev (cid:1)(cid:0)Erev d∆ λli,mj C (Sl,i, Sm,j, ∆)(cid:101)C (κ, κ, ∆) , where l sums over all background spike trains Sl,i pro- jecting to neuron i and m sums over all background spike trains Sm,j projecting to neuron j. (cid:101)C (κ, κ, ∆) is the unnormalized autocorrelation function of the post- synaptic potential i.e., (cid:101)C (κ, κ, ∆) = (cid:104)κ(t)κ(t + ∆)(cid:105), and C (Sl,i, Sm,j, ∆) the cross-correlation function of is given by the background inputs. λli,mj λli,mj = (cid:112)Var (Sl,i) Var (Sm,j). The background cross- correlationisgatedintothecross-correlationofFMPsby thenatureoftherespectivesynapticconnections:ifthetwo neuronsconnecttothecross-correlatedinputsbysynapses of different type (one excitatory, one inhibitory), the sign of the CC is switched (Fig. 2F). However, individual con- tributions to the FMP CC also depend on the difference ofthemeanfreemembranepotentialandthereversalpo- tentials,sothegatingofcross-correlationsisnotsymmet- ricforexcitatoryandinhibitorysynapses.Nevertheless,if (PSP) kernel κ, jm − µj il − µi (cid:88) ρ(ufree wilwjm , ufree ) ∝ (cid:90) l,m · j i 4 τref Figure2:EffectofcorrelatedbackgroundonSSNdynamicsandcompensationthroughreparametrization.(A)FeedforwardreplacementofPoissonnoise byspikingactivityfromotherSSNs.Inthisillustration,theprincipalSSNconsistsofthreeneuronsreceivingbackgroundinputonlyfromotherfunctional SSNsthatsamplefromtheirownpredeterminedtargetdistribution.Forclarity,onlytwooutofatotalof[260,50,34](toptobottomin(B))background SSNsperneuronareshownhere.Bymodifyingthebackgroundconnectivity(grayandbluearrows)theamountofcross-correlationinthebackground inputcanbecontrolled.Atthisstage,thebackgroundSSNsstillreceivePoissoninput(redboxes).(B)Byappropriateparametrizationofthebackground SSNs,weadjustthemeanspikefrequencyofthebackgroundneurons(blue)tostudytheeffectofbackgroundautocorrelations C(Sx, Sx, ∆).Higherfiring probabilitiesincreasethechanceofevokingbursts,whichinducebackgroundautocorrelationsfortheneuronsintheprincipalSSNatmultiplesof τref (darkblue:simulationresults;lightblue: ek ln ¯p with k = ∆ ,seeEq.(10)).(C)BackgroundautocorrelationnarrowstheFMPdistributionofneurons in the principal SSN: simulation (blue bars) and the theoretical prediction (Eq. (11), blue line) vs. background Poisson noise of the same rate (gray). Background intensities correspond to (B). (D) Single-neuron activation functions corresponding to (B,C) and the theoretical prediction (Eq. (12), blue line).Forautocorrelatednoise,theslopeoftheresponsecurvechanges,buttheinflectionpoint(with p(z = 1) = 0.5)isconserved.(E)Kullback-Leibler (cid:0)pnet (cid:107) ptarget(cid:1)(medianandrangebetweenthefirstandthirdquartile)forthethreecasesshownin(B,C,D)aftersamplingfrom50 divergence DKL differenttargetdistributionswith10differentrandomseedsforthe3-neuronnetworkdepictedin(A).Appropriatereparametrizationcanfullycancelout theeffectofbackgroundautocorrelations(blue).Theaccordingresultswithoutreparametrization(gray)andwithPoissoninput(red)arealsoshown.(F)A pairofinterconnectedneuronsinabackgroundSSNgeneratescorrelatednoise,asgivenbyEq.(13).Theeffectofcross-correlatedbackgroundonapair oftargetneuronsdependsonthenatureofsynapticprojectionsfromthebackgroundtotheprincipalSSN.Here,wedepictthecasewheretheirinteraction isexcitatory;theinhibitorycaseisamirrorimagethereof.Left:Ifforwardprojectionsareofthesametype,postsynapticpotentialswillbepositively correlated.Middle:Differentsynapsetypesintheforwardprojectiononlychangethesignofthepostsynaptic potentialcorrelations.Right:Formany backgroundinputswithmixedconnectivitypatterns,correlationscanaverageouttozeroevenwhenallinputcorrelationshavethesamesign.(G)Same experimentasin(E),withbackgroundconnectionstatisticsadjustedtocompensateforinputcross-correlations.Theuncompensatedcasesfrom(F,left) and(F,middle)areshowningray.(H)Correlation-cancellingreparametrizationintheprincipalSSN.Bytransformingthestatespacefrom z ∈ {0, 1}nto z(cid:48) ∈ {−1, 1},inputcorrelationsattainthesamefunctionaleffectassynapticweights(Eq.(15));simulationresultsgivenasreddots,linearfitasredline. Weightrescalingfollowedbyatransformationbackintothe z ∈ {0, 1}nstatespace,showningreen(whichaffectsbothweightsandbiases)cantherefore alleviatetheeffectsofcorrelatedbackground.(I)Similarexperimentasin(E)foranetworkwithtenneurons,withparametersadjustedtocompensatefor inputcross-correlations.Asinthecaseofautocorrelatedbackground,cross-correlationscanbecancelledoutbyappropriatereparametrization. cialsetting,furtheranalysisisneededtoassessitscompat- theconnectivitystatistics(in-degreeandsynapticweights) from the background sources to an SSN are chosen ap- ibilitywiththecorticalconnectomewithrespecttoconnec- tivity statistics or synaptic weight distributions. However, propriatelyandenoughpresynapticpartnersareavailable, even if cortical architecture prevents a clean implementa- thetotalpairwisecross-correlationbetweenneuronsinan tionofthisdecorrelationmechanism,SSNscanthemselves SSNcancelsoutonaverage,leavingthesamplingperfor- manceunimpaired(Fig.2G).Notethatthiswayofreducing compensate for residual background cross-correlations by modifying their parameters, similar to the autocorrelation cross-correlationsisindependentoftheunderlyingweight distributionofthenetworksprovidingthebackground;the compensationdiscussedabove. required cross-wiring of functional networks could there- Todemonstratethisability,weneedtoswitchfromthe natural state space of neurons z ∈ {0, 1}N to the more fore, in principle, be encoded genetically and does not needtobelearned.Furthermore,averysimplecross-wiring symmetricspace z(cid:48) ∈ {−1, 1}N.2Byrequiring p(z(cid:48)) rule, i.e., independently and randomly determined connec- to conserve state probabilities (and thereby also correla- tions,alreadysufficestoaccomplishlowbackgroundcross- correlations and therefore reach a good sampling perfor- 2The z = 0stateforasilentneuronisarguablymorenatural,because ithasnoeffectonitspostsynapticpartnersduringthisstate.Incontrast, mance.Whereasthismethodisguaranteedtoworkinanartifi- z ∈ {−1, 1}would,forexample,implyefferentexcitationuponspikingand constantefferentinhibitionotherwise. ! = p(z) wpre ij 5 exc.inh.randomAF0200400600800simulation time [ms]0.00.51.0v0p(z=1)=0.10.00.51.0(Sx,Sx,)p(z=1)=0.605101520 [ref]0.00.51.0p(z=1)=0.902004006008000200400600800656361u [mV]103102101100101100DKL(pnetptarget)wpreij>0100103105simulation time [ms]103102101100101100wpreij<00.00.10.20.3CC102101IGD0.00.20.4W000.20.40.6CCg1CC=g(W0)H103102101100101100DKL(pnetptarget)103102101100101100100103105simulation time [ms]103102101100101100C-60-56-52-48u [mV]0.00.20.40.00.20.40.00.20.4p(u)EB0.00.51.00.00.51.0p(z=1)-56-54-52-50 [mV]0.00.51.0wpreij>0ρ(ui,uj)>0wpreij>0ρ(ui,uj)<0wpreij>0ρ(ui,uj)≈0z1z2z3 1 4 1 2 1 4 i g1 coliW . W (cid:48) = W and b(cid:48) = b + (cid:88) ρ(Si, Sj) − g0 , w(cid:48)ij = g−1[ρ(Si, Sj)] ≈ tions), the desired change of state variables z(cid:48) = 2z − 1 canbeachievedwithalinearparametertransformation(see Eqs.(66)and(67)): (14) Inthe {−1, 1}N statespace,bothsynapticconnections w(cid:48)ij andbackgroundcross-correlations ρ(Si, Sj)shiftprobabil- ity mass between mixed states (zi, zj) = ±(1,−1) and aligned states (zi, zj) = ±(1, 1) (see Supporting informa- tion,Fig.S1).Therefore,byadjusting band W,itispos- sible to find a W (cid:48) (Fig. 2H) that precisely conserves the desiredcorrelationstructurebetweenneurons: (15) withconstants g0and g1(Fig.2I).Therefore,whenanSSN learns a target distribution from data, background cross- correlationsareequivalenttoanoffsetintheinitialnetwork parametersandareautomaticallycompensatedduringtrain- ing.Fornow,wecanconcludethattheactivityofSSNscon- stitutesasufficientsourceofstochasticityforotherSSNs, since all effects that follow from replacing Poisson noise inanSSNwithfunctionaloutputfromotherSSNs(which at this point still receive explicit noise) can be compen- satedbyappropriateparameteradjustments.Theseareim- portantpreliminaryconclusionsforthenextsections,where we show how all noise can be eliminated in an ensemble of interconnected SSNs endowed with synaptic plasticity withoutsignificantpenaltytotheirrespectivefunctionalper- formance. 3.3 Samplingwithoutexplicitnoiseinlargeen- Weinitializedanensembleof1006-neuronSSNswithan inter-networkconnectivityof  = 0.1andrandomsynaptic weights.Asopposedtothepreviousexperiments,noneofthe neuronsintheensemblereceiveexplicitPoissoninputand theactivityoftheensembleitselfactsasasourceofstochas- ticityinstead,asdepictedinFig.1D.Noexternalinputis neededtokick-startnetworkactivity,assomeneuronsspike spontaneously,duetotherandominitializationofparame- ters(seeFig.3A).Theexistenceofinhibitoryweightsdis- ruptstheinitialregularity,initiatingthesamplingprocess. Ongoinglearning(Equations(8)and(9))shapesthesam- pleddistributionstowardstheirrespectivetargets(Fig.3B), theparametersofwhichweredrawnrandomly(seeSupport- inginformation).Ourensembleachievedasamplingperfor- mance(median DKL)of 1.06+0.27 −0.40×10−3,whichissimilarto themedianperformanceofanidealizedsetup(independent, Poisson-drivenSSNsasin[44])of 1.05+0.15 −0.35 × 10−3(errors aregivenbythefirstandthirdquartile).Toputtheabove sembles strate DKL values in perspective, we compare the sampled and targetdistributionsofoneoftheSSNsintheensembleat variousstagesoflearning(Fig.3C).Thus,despitethefully deterministic nature of the system, the network dynamics and achieved performance after training is essentially in- distinguishable from that of networks harnessing explicit noisefortherepresentationofprobability.Insteadoftrain- ingensembles,theycanalsobesetupbytranslatingthe parametersofthetargetdistributionstoneurosynapticpa- rametersdirectly,asdiscussedintheprevioussection(see Supportinginformation,Fig.S2). 3.4 Implementation on a neuromorphic sub- To test the robustness of our results, we studied an im- plementationofnoise-freesamplingonanartificialneural substrate,whichincorporatesunreliablecomponentsandis thereforesignificantlymoredifficulttocontrol.Forthis,we used the BrainScaleS system [58], a mixed-signal neuro- morphicplatformwithanalogneurosynapticdynamicsand digitalinter-neuroncommunication(Fig.3D,seealsoSup- portinginformation,Fig.S3).Amajoradvantageofthisim- plementationistheemulationspeedupof 104withrespectto biologicalreal-time;however,forclarity,weshallcontinue usingbiologicaltimeunitsinsteadofactualemulationtime. The additional challenge for our neuronal ensemble is tocopewiththenaturalvariabilityofthesubstrate,caused mainlybyfixed-patternnoise,orwithotherlimitationssuch as a finite weight resolution (4 bits) or spike loss, which canallbesubstantial[59,60].Itisimportanttonotethat theabilitytofunctionwhenembeddedinanimperfectsub- strate with significant deviations from an idealized model represents a necessary prerequisite for viable theories of biologicalneuralfunction. We emulated an ensemble of 15 4-neuron SSNs, with an inter-SSN connectivity of  = 0.2 and with randomly drawntargetdistributions(seeSupportinginformation).The biases were provided by additional bias neurons and ad- justed during learning via the synaptic weights between biasandsamplingneurons,alongwiththesynapseswithin the SSNs, using the same learning rule as before (Equa- tions (8) and (9)). After 200 training steps, the ensemble −1.15 · 10−2 (errors given reached a median DKL of 3.99+1.27 by the distance to the first and third quartile) compared −0.55 before training (Fig. 3E). As a point of refer- to 1.18+0.47 ence,wealsoconsideredtheidealizedcasebytrainingthe samesetofSSNswithoutinterconnectionsandwithevery neuronreceivingexternalPoissonnoisegeneratedfromthe hostcomputer,reachinga DKLof 2.49+3.18 This relatively small performance loss of the noise-free ensemblecomparedtotheidealcaseconfirmsthetheoreti- calpredictionsandsimulationresults.Importantly,thiswas achievedwithonlyarathersmallensemble,demonstrating thatlargenumbersofneuronsarenotneededforrealizing −0.71 · 10−2. 6 Figure3:Samplingwithoutexplicitnoisefromasetofpredefinedtargetdistributionsinsoftware(A-C)andonaneuromorphicsubstrate(D-G).(A)(top) Temporalevolutionofspikingactivityinanensembleof100interconnected6-neuronSSNswithnosourceofexplicitnoise.Aninitialburstofregular activitycausedbyneuronswithastrongenoughpositivebiasquicklytransitionstoasynchronousirregularactivityduetoinhibitorysynapses.(bottom) Distribution of mean neuronal firing rates of the ensembles shown in (B,C) after training. (B) Median sampling quality of the above ensemble during learning,foratestsamplingperiodof 106ms.Attheendofthelearningphase,thesamplingqualityofindividualnetworksintheensemble(blue)ison parwiththeoneobtainedinthetheoreticallyidealcaseofindependentnetworkswithPoissonbackground(black).Errorbarsgivenover5simulationruns withdifferentrandomseeds.(C)Illustrationofasingletargetdistribution(magenta)andcorrespondingsampleddistribution(blue)ofanetworkinthe ensembleatseveralstagesofthelearningprocess(dashedgreenlinesin(B)).(D)PhotographofawaferfromtheBrainScaleSneuromorphicsystemused in(E),(F)and(G)beforepost-processing(i.e.,addingadditionalstructureslikebusesontop),whichwouldmasktheunderlyingmodularstructure.Blue: exemplarymembranetraceofananalogneuronreceivingPoissonnoise.(E)Performanceofanensembleconsistingof154-neuronSSNswithnoexternal noiseduringlearningontheneuromorphicsubstrate,showninlightblueforeachSSNandwiththemedianshownindarkblue.Thelargefluctuations comparedto(B)areasignatureofthenaturalvariabilityofthesubstrate'sanalogcomponents.Thedashedbluelinerepresentsthebestachievedmedian (cid:0)pnet (cid:107) ptarget(cid:1) = 3.99 × 10−2.Forcomparison,wealsoplottheoptimalmedianperformanceforthetheoreticallyidealcaseof performanceat DKL (cid:0)pnet (cid:107) ptarget(cid:1) = 2.49 × 10−2 (dashedblackline).(F)Left: independent,Poisson-drivenSSNsemulatedonthesamesubstrate,whichliesat DKL DemonstrationofsamplingintheneuromorphicensembleofSSNsafter200trainingsteps.Individualnetworksinlightblue,medianperformanceindark blue.Dashedblueline:medianperformancebeforetraining.Dashedblackline:medianperformanceofidealnetworks,asin(E).Right:Bestachieved performance,after 100 sofbiotime(10 msofhardwaretime)forallSSNsintheensembledepictedasbluedots(sortedfromlowesttohighest DKL). Forcomparison,thesameisplottedasblackcrossesfortheiridealcounterparts.(G)Sampled(blue)andtarget(magenta)distributionsoffourofthe15 SSNs.Theselectionismarkedin(F)withgreentriangles(left)andverticalgreendashedlines(right).Sincewemadenoparticularselectionofhardware neuronsaccordingtotheirbehavior,hardwaredefectshaveasignificantimpactonasmallsubsetoftheSSNs.Despitetheseimperfections,amajorityof SSNsperformclosetothebestvaluepermittedbythelimitedweightresolution(4bits)ofthesubstrate. thiscomputationalparadigm. InFig.3FandVideoS1,weshowthesamplingdynamics ofallemulatedSSNsafterlearning.WhilemostSSNsare able to approximate their target distributions well, some sampled distributions are significantly skewed (Fig. 3G). Thisiscausedbyasmallsubsetofdysfunctionalneurons, whichwehavenotdiscardedbeforehand,inordertoavoid animplausiblyfine-tuneduse-caseoftheneuromorphicsub- strate. These effects become less significant in larger net- works trained on data instead of predefined distributions, wherelearningcannaturallycopewithsuchoutliersbyas- signingthemsmalleroutputweights.Nevertheless,thesere- sultsdemonstratethefeasibilityofself-sustainedBayesian computationthroughsamplinginphysicalneuralsubstrates, without the need for any source of explicit noise. Impor- tantly,andincontrasttootherapproaches[61],everyneu- ronintheensembleplaysafunctionalrole,withnoneuronal real-estate being dedicated to the production of (pseudo- )randomness. 3.5 EnsemblesofhierarchicalSSNs When endowed with appropriate learning rules, hierarchi- cal spiking networks can be efficiently trained on high- dimensionalvisualdata[62,60,47,63 -- 65].Suchhierarchi- calnetworksarecharacterizedbythepresenceofseveral layers,withconnectionsbetweenconsecutivelayers,butno lateralconnectionswithinthelayersthemselves.Whenboth feedforwardandfeedbackconnectionsarepresent,suchnet- worksareabletobothclassifyandgenerateimagesthatare similartothoseusedduringtraining. Inthesenetworks,informationprocessinginbothdirec- tions is Bayesian in nature. Bottom-up propagation of in- formationenablesanestimationoftheconditionalprobabil- ityofaparticularlabeltofittheinputdata.Additionally, top-downpropagationofneuralactivityallowsgeneratinga subsetofpatternsinthevisiblelayerconditionedonincom- pleteorpartiallyoccludedvisualstimulus.Whennoinput is presented, such networks will produce patterns similar 7 100101102103training step103102101100101100DKL(pnetptarget)Bz101102103104p(z)2000 steps, DKL9.31104100101102training steps102101100DKL(pnetptarget)EA101103105biol. time [ms]102101100DKL(pnetptarget)FC00.20.40.60.81.0mean rate r [1ref]012p(r)z101102103104p(z)100 steps, DKL0.12z101102103104p(z)5 steps, DKL2.33Dz103102101p(z)DKL 0.017GzDKL 0.0360100200sim. time [ms]neuronszDKL 0.419151015networkszDKL 0.048 Figure4:Bayesianinferenceonvisualinput.(A)IllustrationoftheconnectivitybetweentwohierarchicalSSNsinthesimulatedensemble.EachSSN hadavisiblelayerv,ahiddenhandalabellayerl.NeuronsinthesamelayerofanSSNwerenotinterconnected.EachneuroninanSSNreceived onlyactivityfromthehiddenlayersofotherSSNsasbackground(nosourcesofexplicitnoise).(B)AnensembleoffoursuchSSNs(red)wastrainedto performgenerativeanddiscriminativetasksonvisualdatafromtheEMNISTdataset.WeusedtheclassificationrateofrestrictedBoltzmannmachines trainedwiththesamehyperparametersasabenchmark(blue).Errorbarsaregiven(onblue)over10testrunsand(onred)over10ensemblerealizations withdifferentrandomseeds.(C)IllustrationofascenariowhereoneofthefourSSNs(redboxes)receivedvisualinputforclassification(B).Atthesame time,theotherSSNscontinuouslygeneratedimagesfromtheirrespectivelearneddistributions.(D)Patterngenerationandmixingduringunconstrained dreaming. Here, we show the activity of the visible layer of all four networks from (B), each spanning three rows. Time evolves from left to right. For furtherillustrationsofthesamplingprocessintheensembleofhierarchicalSSNs,seeSupportinginformation,Figs.S4andS5 andVideoS2.(E)Pattern completionandrivalryfortwoinstancesofincompletevisualstimulus.Thestimulusconsistedofthetoprightandbottomrightquadrantofthevisible layer,respectively.Inthefirstrun,weclampedthetoparcofa"B"compatiblewitheithera"B"oran"R"(topthreerows,red),inthesecondrunwe chosethebottomlineofan"L"compatiblewithan"L",an"E",a"Z"ora"C"(bottomthreerows,red).AnensembleofSSNsperformsBayesianinference byimplicitlyevaluatingtheconditionaldistributionoftheunstimulatedvisibleneurons,whichmanifestsitselfhereassamplingfromallimageclasses compatiblewiththeambiguousstimulus(seealsoSupportinginformation,Fig.S6). −3.0%witherrorsgivenbythedistancetothefirst of 91.5+3.6 tothoseenforcedduringtraining("dreaming").Ingeneral, andthirdquartile,whichisclosetothe 94.0+2.1 −1.5%achieved the exploration of a multimodal solution space in genera- by the idealized reference setup provided by the abstract tive models is facilitated by some noise-generating mech- Boltzmannmachines(Fig.4B).Atthesametime,allother anism. We demonstrate how even a small interconnected set of hierarchical SSNs can perform these computations SSNsremainedcapableofgeneratingrecognizableimages (Fig. 4C). It is expected that direct training and a larger self-sufficiently,withoutanysourceofexplicitnoise. numberofSSNsintheensemblewouldfurtherimprovethe Weusedanensembleoffour3-layerhierarchicalSSNs results, but a functioning translation from the abstract to trained on a subset of the EMNIST dataset [66], an ex- thebiologicaldomainalreadyunderpinsthesoundnessof tendedversionofthewidelyusedMNISTdataset[67]that theunderlyingtheory. includesdigitsaswellascapitalandlower-caseletters.All Withoutvisualstimulus,allSSNssampledfreely,gener- SSNshadthesamestructure,with784visibleunits,200 ating images similar to those on which they were trained hidden units and 5 label units (Fig. 4A). To emulate the (Fig.4D).Withoutanysourceofexplicitnoise,theSSNs presenceofnetworkswithdifferentfunctionality,wetrained were capable to mix between the relevant modes (images eachofthemonaseparatesubsetofthedata.(Tocombine belongingtoallclasses)oftheirrespectiveunderlyingdis- samplinginspacewithsamplingintime,multiplenetworks tributions,whichisahallmarkofagoodgenerativemodel. can also be trained on the same data, see Supporting in- Wefurtherextendedtheseresultstoanensembletrained formationFig.S5 andVideoS2.)Sincetrainingthespik- on the full MNIST dataset, reaching a similar generative ingensembledirectlywascomputationallyprohibitive,we performance for all networks (see Supporting information trainedfourBoltzmannmachinesontherespectivedatasets Fig.S5 andVideoS2). andthentranslatedtheresultingparameterstoneurosynap- TotestthepatterncompletioncapabilitiesoftheSSNs ticparametersoftheensembleusingthemethodsdescribed intheensemble,westimulatedthemwithincompleteand earlierinthemanuscript(seealsoSupportinginformation, ambiguous visual data (Fig. 4E). Under these conditions, Fig.S2). SSNsonlyproducedimagescompatiblewiththestimulus, TotestthediscriminativepropertiesoftheSSNsinthe alternating between different image classes, in a display ensemble,onewasstimulatedwithvisualinput,whilethe ofpatternrivalry.Asinthecaseoffreedreaming,thekey remainingthreewerelefttofreelysamplefromtheirunder- mechanismfacilitatingthisformofexplorationwasprovided lyingdistribution.Wemeasuredamedianclassificationrate 8 CBEACELZ13579Aa15YW0.40.60.81.0BRTXVDvhl(visible)(hidden)(label) bythefunctionalactivityofotherneuronsintheensemble. 4 Discussion Based on our findings, we argue that sampling-based Bayesian computation can be implemented in fully deter- ministic ensembles of spiking networks without requiring anyexplicitnoise-generatingmechanism.Ourapproachhas afirmtheoreticalfoundationinthetheoryofsamplingspik- ing neural networks, upon which we formulate a rigorous analysisofnetworkdynamicsandlearninginthepresence orabsenceofnoise. While in biology various explicit sources of noise exist [68 -- 70],theseformsofstochasticityareeithertooweak(in caseofionchannels)ortoohigh-dimensionalforefficient exploration(inthecaseofstochasticsynaptictransmission, asusedfor,e.g.,reinforcementlearning[71]).Furthermore, arigorousmathematicalframeworkforneuralsamplingwith stochasticsynapsesisstilllacking.Ontheotherhand,in thecaseofpopulationcodes,neuronalpopulationnoisecan be highly correlated, affecting information processing by, e.g.,inducingsystematicsamplingbiases[33]. Inourproposedframework,eachnetworkinanensemble playsadualrole:whilefulfillingitsassignedfunctionwithin itshomesubnetwork,italsoprovidesitspeerswiththespik- ingbackgroundnecessaryforstochasticsearchwithintheir respective solution spaces. This enables a self-consistent and parsimonious implementation of neural sampling, by allowing all neurons to take on a functional role and not dedicatinganyresourcespurelytotheproductionofback- groundstochasticity.Theunderlyingidealiesinadapting neuro-synapticparametersby(contrastive)Hebbianlearn- ingtocompensateforauto-andcross-correlationsinduced byinteractionsbetweenthefunctionalnetworksintheen- semble. Importantly, we show that this does not rely on thepresenceofalargenumberofindependentpresynaptic partners for each neuron, as often assumed by models of corticalcomputationthatusePoissonnoise(see,e.g.,[72]). Instead,onlyasmallnumberofensemblesisnecessaryto implementnoise-freeBayesiansampling.Thisbecomespar- ticularlyrelevantforthedevelopmentofneuromorphicplat- forms by eliminating the computational footprint imposed bythegenerationanddistributionofexplicitnoise,thereby reducingpowerconsumptionandbandwidthconstraints. Forsimplicity,wechosenetworksofsimilarsizeinour simulations.However,thepresentedresultsarenotcontin- gentonnetworksizesintheensembleandlargelyindepen- dentoftheparticularfunctionality(underlyingdistribution) ofeachSSN.Theirapplicabilitytoscenarioswherediffer- ent SSNs learn to represent different data is particularly relevant for cortical computation, where weakly intercon- nected areas or modules are responsible for distinct func- tions [73 -- 77]. Importantly, these ideas scale naturally to largerensemblesandlargerSSNs.Sinceeachneurononly needsasmallnumberofpresynapticpartnersfromtheen- 9 semble,largernetworksleadtoasparserinterconnectivity betweenSSNsintheensembleandhencesoftenstructural constraints.Preliminarysimulationsshowthattheprinciple ofusingfunctionaloutputasnoisecanevenbeappliedto connectionswithinasingleSSN,eliminatingtheartificial separationbetweennetworkandensembleconnections(see Fig.S7andVideoS3intheSupportinginformation). Eventhoughwehaveusedasimplifiedneuronmodelin our simulations to reduce computation time and facilitate the mathematical analysis, we expect the core underlying principlestogeneralize.Thisisevidencedbyourresultson neuromorphic hardware, where the dynamics of individual neurons and synapses differ significantly from the mathe- maticalmodel.Suchanabilitytocomputewithunreliable componentsrepresentsaparticularlyappealingfeaturein the context of both biology and emerging nanoscale tech- nologies. Finally,thesuggestednoise-freeBayesianbrainrecon- cilesthedebateonspatialversustemporalsampling[78,29]. Infact,thenetworksofspikingneuronsthatprovideeach other with virtual noise may be arranged in parallel sen- sorystreams.Anambiguousstimuluswilltriggerdifferent representationsoneachlevelofthesestreams,formingahi- erarchyofprobabilisticpopulationcodes.Whilethesepop- ulationcodeslearntocoverthefullsensorydistributionin space,theywillalsogeneratesamplesofthesensorydistri- butionintime(seeFig.S5intheSupportinginformation). Attentionmayselectthemostlikelyrepresentation,while suppressingtherepresentationsintheotherstreams.Anal- ogously, possible actions may be represented in parallel motor streams during planning and a motor decision may selecttheonetobeperformed.Whenrecordinginpremotor cortex,suchaselectioncausesanoisereduction[79],that wesuggestiseffectivelythesignatureofchoosingthemost probableactioninaBayesiansense. 5 Conclusion From a generic Bayesian perspective, cortical networks canbeviewedasgeneratorsoftargetdistributions.Toen- ablesuchcomputation,modelsassumeneuronstopossess sourcesofperfect,well-behavednoise -- anassumptionthat is both impractical and at odds with biology. We showed howlocalplasticityinanensembleofspikingnetworksal- lowsthemtoco-shapetheiractivitytowardsasetofwell- definedtargets,whilereciprocallyusingtheverysameac- tivity as a source of (pseudo-)stochasticity. This enables purelydeterministicnetworkstosimultaneouslylearnava- rietyoftasks,completelyremovingtheneedfortruerandom- ness. While reconciling the sampling hypothesis with the deterministic nature of single neurons, this also offers an efficientblueprintforin-silicoimplementationsofsampling- basedinference. (cid:115) (cid:18) 6 Calculations 6.1 Freemembranepotentialdistributionwith colorednoise In the high-conductance state (HCS), it can be shown thatthetemporalevolutionofthefreemembranepotential (FMP)ofanLIFneuronstimulatedbybalancedPoissonin- putsisequivalenttoanOrnstein-Uhlenbeck(OU)process withthefollowingGreen'sfunction[80]: 1 (cid:19) . , , , k − θ = µ = νkw2 σ2 = k∈{e,i} k∈{e,i} k∈{e,i} · exp k τ syn νkwkErev 1 τ syn , 1 − e−2θt f (u, tu0) = with k − µ)2τ syn (cid:104)gtot(cid:105) = gl + (u − µ + (µ − u0)e−θt)2 (cid:104)gtot(cid:105) k(Erev (cid:104)gtot(cid:105)2 wkνkτ syn 2πσ2(1 − e−2θt) 1 2σ2 glEl +(cid:80) (cid:80) (cid:88) (16) (17a) (17b) (17c) (17d) where νk arethenoisefrequencies, wk thenoiseweights usedinprevious andwedroppedtheindexnotation ufree sectionsforconvenience.ThestationaryFMPdistribution isthengivenbyaGaussian[26,80]: (cid:19) (cid:18) (18) Replacingthewhitenoise η(t)intheOUprocess,defined by (cid:104)η(cid:105) = const.and (cid:104)η(t)η(t(cid:48))(cid:105) = νδ(t − t(cid:48)) + ν2[26],with (Gaussian)colorednoise ηc,definedby (cid:104)ηc(cid:105) = const.and (cid:104)ηc(t)ηc(t(cid:48))(cid:105) = γ(t − t(cid:48))[81]where γ(t − t(cid:48))isafunction thatdoesnotvanishfor t− t(cid:48) (cid:54)= 0,thestationarysolutionof theFMPdistributionisstillgivenbyaGaussianwithmean µ(cid:48) andwidth σ(cid:48) [81,82].Sincethenoisecorrelationsonly appearwhencalculatinghigher-ordermomentsoftheFMP, themeanvalueoftheFMPdistributionremainsunchanged (cid:1)2 µ(cid:48) = µ.However,thevariance σ(cid:48)2 = (cid:104) (cid:105)of thestationaryFMPdistributionchangesduetothecorre- lations,asdiscussedinthenextsection. (cid:0)u(t) − (cid:104)u(t)(cid:105) 1 2πσ2 exp (u − µ)2 (cid:114) f (u) = 2σ2 − k . 10 tion 6.2 Widthoffreemembranepotentialdistribu- IntheHCS,theFMPcanbeapproximatedanalyticallyas [80] (cid:88) (cid:88) (cid:19) i , · (cid:18) (cid:19)(cid:21) , (cid:18) (cid:20) · Λk − − u(t) = u0 + ΛkΘ(t − ts) exp exp spikes s u0 = Λk = (cid:104)gtot(cid:105) − exp − exp u(t) = u0 + (cid:104)τeff(cid:105) = t − t(cid:48) (cid:104)τeff(cid:105) t − ts (cid:104)τeff(cid:105) with k − (cid:104)τeff(cid:105) k∈{e,i} t − ts − τ syn k k∈{e,i} t − t(cid:48) − τ syn k = u0 + Λe (cid:88) (cid:19) (cid:0)Se ∗ κe (cid:18) glEl + ((cid:104)gtot(cid:105) − gl)µ (cid:1) , τ syn k wk (cid:104)gtot(cid:105) Cm (cid:104)gtot(cid:105) (19) (20a) (cid:0)Erev k − µ(cid:1) (20b) (cid:0)τ syn (20c) By explicitly writing the excitatory and inhibitory noise spike trains as Se/i(t(cid:48)) = (cid:80) spikes s δ(t(cid:48) − ts), this can be rewrittento (cid:90) (cid:18) (cid:19)(cid:21) (cid:20) (cid:18) dt(cid:48)Sk(t(cid:48))Θ(t − t(cid:48)) (21a) (cid:0)Si ∗ κi (cid:1)(t) (cid:1)(t) + Λi (21b) = u0 +(cid:2)(ΛeSe + ΛiSi) ∗ κ(cid:3)(t) , (21c) where ∗denotestheconvolutionoperatorandwith (cid:20) (cid:19) (cid:18) (cid:19)(cid:21) . (22) . The width of the For simplicity, we assume τ syn FMPdistributioncannowbecalculatedas (cid:1)2 (cid:0)u(t) − (cid:104)u(t)(cid:105) (23a) (cid:2)u0 + (Stot ∗ κ)(t)(cid:3)2 (23b) (cid:2)(Stot ∗ κ)(t)(cid:3)2 (23c) (cid:2)(Stot ∗ κ)(t) − (cid:104)(Stot ∗ κ)(t)(cid:105) (23d) where the average is calculated over t and Stot(t) = ΛeSe(t) + ΛiSi(t). Since the average is an integral over (cid:82) T /2 −T /2(·) dt,wecanusetheidentity t,i.e. (cid:104)(·)(cid:105) → lim (cid:82) (f ∗ g)(t) dt = ((cid:82) f (t) dt)((cid:82) g(t) dt), that is (cid:104)f ∗ g(cid:105) = (cid:82) g(t) dt = (cid:104)f(cid:105) ∗ ginthelimitof T → ∞,toarriveat thefollowingsolution: (cid:0)u(t) − (cid:104)u(t)(cid:105) (cid:1)2 (cid:2)(Stot(t) − (cid:104)Stot(t)(cid:105)) ∗ κ(t)(cid:3)2 (24) (cid:105) = (cid:104)u(t)2(cid:105) − (cid:104)u(t)(cid:105)2 (cid:3)2 (cid:105) − (cid:104)(Stot ∗ κ)(t)(cid:105)2 (cid:105) , (cid:105) − (cid:104)u0 + (Stot ∗ κ)(t)(cid:105)2 (cid:104) = (cid:104) = (cid:104) = (cid:104) e = τ syn κe/i(t) = Θ(t) (cid:104) = (cid:104) − exp (cid:104)τeff(cid:105) T→∞ τ syn e/i (cid:104)f(cid:105) (cid:105) . exp − − t t 1 T . (cid:105) 1 1 (cid:105) lim (cid:105) , (cid:104)¯u(t)¯u(t + ∆)(cid:105) = (cid:104) (cid:104) = lim T→∞ = lim T→∞ = lim T→∞ More generally, we obtain with a similar calculation the autocorrelationfunction(ACF)oftheFMP: (cid:0)( ¯Stot ∗ κ)(t)(cid:1)(cid:0)( ¯Stot ∗ κ)(t + ∆)(cid:1) (25) with ¯x(t) = x(t) − (cid:104)x(t)(cid:105) and by using (cid:104)¯u(t)¯u(t + ∆)(cid:105) = (cid:104)u(t)u(t + ∆)(cid:105) − (cid:104)u(t)(cid:105)2.Thiscanbefurthersimplifiedby applying the Wiener -- Khintchine theorem [83, 84], which T→∞(cid:104)x(t)x(t + ∆)(cid:105)T = F−1(cid:0) F(x)2)(cid:1)(∆) states that −T /2(·) dt (due to (cid:82) x(t)x(t + ∆) dt = (cid:82) T /2 with (cid:104)(·)(cid:105)T → (cid:0)x(t) ∗ x(−t)(cid:1)(∆)). Thus, for the limit T → ∞, we can rewritethisas (cid:0)( ¯Stot ∗ κ)(t)(cid:1)(cid:0)( ¯Stot ∗ κ)(t + ∆)(cid:1) T F−1(cid:0) F( ¯Stot ∗ κ)2(cid:1)(∆) (26a) F( ¯Stot)F(κ)2(cid:1)(∆) T F−1(cid:0) (26b) (cid:18) F−1(cid:0) F( ¯Stot)2(cid:1) F(κ)2(cid:1)(cid:19) ∗ F−1(cid:0) (26c) andbyapplyingtheWiener -- Khintchinetheoremagainin reverse (cid:104)¯u(t)¯u(t + ∆)(cid:105) =(cid:0) lim (27) where the variance of the FMP distribution is given for ∆ = 0. Thus, the unnormalized ACF of the FMP can be calculatedbyconvolvingtheunnormalizedACFoftheback- groundspiketrains(Stot)andthePSPshape(κ).Incase ofindependentexcitatoryandinhibitoryPoissonnoise(i.e., (cid:104) ¯S(t) ¯S(t(cid:48))(cid:105) = νδ(t − t(cid:48))),weget (cid:2) ¯Stot(t)(cid:3)(cid:2) ¯Stot(t + ∆(cid:48))(cid:3) (28a) e (cid:104) ¯Se(t) ¯Se(t + ∆(cid:48))(cid:105) (cid:105) = Λ2 (cid:88) (28b) i (cid:104) ¯Si(t) ¯Si(t + ∆(cid:48))(cid:105) + Λ2 (cid:2) ¯Stot(t)(cid:3)(cid:2) ¯Stot(t + ∆(cid:48))(cid:3) (cid:2)κ(t)(cid:3)(cid:2)κ(t + ∆(cid:48))(cid:3) (cid:1)(∆) , 1 T (cid:104) T→∞ (cid:105)∞ (∆) , ∗ (cid:104) (cid:105)T 1 T kνkδ(∆(cid:48)) Λ2 = (cid:104) andtherefore k∈{e,i} Λ2 kνkδ(∆(cid:48)) Var(u) =(cid:0) (cid:88) (cid:2)κ(t)(cid:3)(cid:2)κ(t + ∆(cid:48))(cid:3) (cid:88) (cid:90) ∞ (cid:88) Λ2 kνk (cid:104)κ2(t)(cid:105) ∗ (cid:104) = k∈{e,i} k∈{e,i} = Λ2 kνk κ2(t) dt , 0 k∈{e,i} (cid:1)(∆ = 0) (cid:105)∞ (29a) (29b) (29c) 11 = Λ2 β = kνk k∈{e,i} (cid:80) 0 κ2(t) dt σ2 colored σ2 Poisson (cid:82) dt κ(t)κ(t + ∆) (cid:82) d∆ (cid:104) ¯Stot(t) ¯Stot(t + ∆)(cid:105) · (cid:82) ∞ whichagreeswiththeresultgivenin[80].Ifthenoisespike trains are generated by processes with refractory periods, theabsenceofspikesbetweenrefractoryperiodsleadsto negativecontributionsintheACFofthenoisespiketrains. ThisleadstoareducedvalueofthevarianceoftheFMP andhence,alsotoareducedwidthoftheFMPdistribution. Thefactor √βbywhichthewidthoftheFMPdistribution (Eqn.(11))changesduetotheintroductionofcoloredback- groundnoiseisgivenby (30) . (31) ForthesimplifiedcaseofaPoissonprocesswithrefractory period, one can show that (cid:82) d∆ (cid:104) ¯Stot(t) ¯Stot(t + ∆)(cid:105) has areducedvaluecomparedtoaPoissonprocesswithoutre- fractoryperiod[26],leadingto β ≤ 1.Eventhoughwedo not show this here for neuron-generated spike trains, the twocasesaresimilarenoughthat β ≤ 1canbeassumed toapplyinthiscaseaswell. Inthenextsection,wewillshowthatthefactor βcanbe usedtorescaletheinverseslopeoftheactivationfunction to transform the activation function of a neuron receiving whitenoisetotheactivationfunctionofaneuronreceiving equivalent (in frequency and weights), but colored noise. That is, the rescaling of the FMP distribution width due to the autocorrelated background noise translates into a rescaledinverseslopeoftheactivationfunction. 6.3 ApproximateinverseslopeofLIFactivation As stated earlier, the FMP of an LIF neuron in the HCS isdescribedbyanOUprocesswithaGaussianstationary FMP distribution (both for white and colored background noise).Asafirstapproximation,wecandefinetheactivation function as the probability of the neuron having a FMP abovethreshold(seeEq.(18)) (32) (33) (34) Eventhoughthisisonlyanapproximation(aswearene- glectingtheeffectofthereset),theerrorfunctionisalready similartothelogisticactivationfunctionobservedinsimu- lations[44]. Theinverseslopeofalogisticactivationfunctionisde- (cid:90) ∞ (cid:114) (cid:18) (cid:90) ∞ (cid:1)(cid:19) (cid:18) 1 − erf(cid:0) ϑ − µ function 1 2πσ2 exp p(zi = 1) ≈ (u − µ)2 √2σ ϑ 1 2 f (u)du (cid:19) 2σ2 − du = = ϑ . , . α ϕ = d dµ d dµ α−1 = α−1 = (cid:114) p(zi = 1) 1 2πσ2 , finedattheinflectionpoint,i.e., (cid:19)(cid:12)(cid:12)(cid:12)(cid:12)µ=u0 (cid:18) µ − u0 (35) Bycalculatingtheinverseslopeviatheactivationfunction derivedfromtheFMPdistribution,weget (cid:12)(cid:12)(cid:12)(cid:12)µ=ϑ (36) (37) fromwhichitfollowsthattheinverseslope αisproportional to the width of the FMP distribution σ. Thus, rescaling the variance of the FMP distribution by a factor β leads, approximately, to a rescaling of the inverse slope of the activationfunction α(cid:48) = √βα. 6.4 Originofside-peaksinthenoiseautocor- relationfunction Forhighrates,thespiketraingeneratedbyanLIFneuron intheHCSshowsregularpatternsofinterspikeintervals whichareroughlyequaltotheabsoluterefractoryperiod. Thisoccurs(i)duetotherefractoryperiodintroducingreg- ularityforhigherrates,sinceISI's < τref arenotallowed andthemaximumfiringrateoftheLIFneuronisbounded by 1 ,and(ii)duetoanLIFneurons'stendencytospike consecutivelywhentheeffectivemembranepotential (38) (39) issuprathresholdaftertherefractoryperiod[80].Theprob- abilityofaconsecutivespikeaftertheneuronhasspiked onceattime tisgivenby(undertheassumptionoftheHCS) (40) following an due to the effective membrane potential Ornstein-Uhlenbeck process (whereas the FMP is a low- pass filter thereof, however with a very low time constant τeff),seeEq.(16).Theprobabilitytospikeagainafterthe secondrefractoryperiodisthengivenby (41) p1 = p(spike at t + τref first spike at t) dut+τref f (ut+τref , τref ut = ϑ) , glEl +(cid:80) p2 = p(spike at t + 2τref spike at t + τref , t) τeff u = ueff − u , (cid:90) ∞ gsyn k (t)Erev k∈{e,i} gtot(t) ueff (t) = τref = ϑ k , = ϑ du2du1 f (u2, τref u1)f (u1, τref u0 = ϑ) ϑ du1 f (u1, τref u0 = ϑ) (cid:82) ∞ ϑ (cid:82) ∞ (cid:82) ∞ , 12 pn = p(spike at t + nτref spike at t + (n − 1)τref , . . . , t) with un = ut+nτref,oringeneralafter n − 1spikes (cid:90) ∞ (cid:82) ∞ ϑ = i=1 pi , Pn = f n(un) = n−1(cid:89) (cid:82) ∞ dun−1 f (n−1)(un−1) , ϑ dun−1 f (n−1)(un−1) p(burst of length n) = Pn · (1 − pn) . ϑ dun−1 f (un, τref un−1)f (n−1)(un−1) (42) ,(43) for n > 1and f 1(u1) = f (u1, τref u0 = ϑ).Theprobability toobserve nspikesinsuchasequenceisthengivenby (44) andtheprobabilitytofindaburstoflength n(i.e.,theburst ends) (45) Withthis,onecancalculatetheaveragelengthoftheoccur- ringbursts(cid:80)∞i=1 i· p(burst of length i),fromwhichwecan already see how the occurrence of bursts depends on the meanactivityoftheneuron.Asimplesolutioncanbefound forthespecialcaseof τ syn (cid:28) τref,sincethentheeffective membranepotentialdistributionhasalreadyconvergedto thestationarydistributionaftereveryrefractoryperiod,i.e., f (un, τref un−1) = f (un)andhence (46) forall n.Thus,forthisspecialcasetheaverageburstlength canbeexpressedas ∞(cid:88) (47) (48) Bychangingthemeanmembranepotential(e.g.,byadjust- ingtheleakpotentialoraddinganexternal(bias)current), theprobabilityofconsecutivespikes ¯pcanbedirectlyad- justedandhence,alsotheaveragelengthofbursts.Since these bursts are fixed structures with interspike intervals equaltotherefractoryperiod,theytranslateintoside-peaks atmultiplesoftherefractoryperiodinthespiketrainACF, aswedemonstratebelow. TheACFofthespiketrain Sisgivenby (49) C(S, S, ∆) = (cid:104)StSt+∆(cid:105) − (cid:104)S(cid:105)2 where the first term of the numerator is (cid:104)StSt+∆(cid:105) = (21a) p(spike at t + ∆, spike at t) (notation as in Eqs. pn = p(spike at t + nτref spike at t + (n − 1)τref , . . . , t) i · p(burst of length i) = i · ¯pi−1(1 − ¯p) , i=1 1 1 − ¯p (cid:90) ∞ ∞(cid:88) duf (u) = ¯p Var(S) i=1 = = ϑ . , and(23a)).Thistermcanbeexpressedas Var(S) a can with spike bursts for ∆ < τref, p(spike at t + ∆, spike at t) = p(spike at t + ∆ spike at t) · p(spike at t) , = p(spike at t + ∆ spike at t) · (cid:104)S(cid:105) , (50) (51) where we assumed that the first spike starts the burst at arandomtime t.Therefore,inordertocalculatetheACF, wehavetocalculatetheprobabilitythataspikeoccursat time t + ∆giventhattheneuronspikesattime t.Thishas toincludeeverypossiblecombinationofspikesduringthis interval.Inthefollowing,wearguethatatmultiplesofthe refractoryperiod,themaincontributiontotheACFcomes frombursts. • First, the term p(spike at t + ∆ spike at t)inEq.(50)vanishessincetheneuronis refractoryandcannotspikeduringthisinterval.Thus, the ACF becomes negative as only the term − (cid:104)S(cid:105)2 inEq.(49)remains,wherebothnumeratoranddenom- inatorarepositive. only τref, • For ∆ = occur probability when neuron the p1 = (cid:82) ∞ ϑ dut+τref f (ut+τref , τref ut = ϑ),wherewe assumed for simplicity that the first spike starts the burstspontaneously. • Sincefor τref < ∆ < 2τref,theneurondidnotburst withprobability 1 − p1,itispossibletofindaspikein thisinterval,leadingagaintonegative,butdiminished, valuesintheACF. • For ∆ = 2τref,wenowhavetwowaystoobservespikes at tand t + 2τref:(i)Thespikesarepartofaburstof length 2 or (ii) there was no intermediate spike and thespikeshaveanISIof 2τref.Sinceforlargerrates, having large ISIs that are exact multiples of τref is unlikely,wecanneglectthecontributionof(ii). • Ifwegofurtherto ∆ = nτref,wegetevenmoreaddi- tionaltermsincludingburstsoflength < n.However, these terms can again be neglected as, compared to havingaburstoflength n,itisratherunlikelytoget a burst pattern with missing intermediate spikes, i.e., havingpartialburstswhichareamultipleof τrefapart. for ∆ → ∞, we have (cid:104)StSt+∆(cid:105) − (cid:104)S(cid:105)2 = • Finally, (cid:104)St(cid:105)(cid:104)St+∆(cid:105) − (cid:104)S(cid:105)2 = (cid:104)S(cid:105)(cid:104)S(cid:105) − (cid:104)S(cid:105)2 = 0andtheACF (Eq.(49))vanishes. Consequently,wecanapproximatetheACFatmultiples of the refractory period by calculating the probability of findingaburstof nspikes(Eq.(43)): (52) Pk+1δ(cid:0)[n − k]τref C(S, S, nτref ) ≈ ∞(cid:88) (cid:1) , k=1 k=1 = k=1 ∞(cid:88) ∞(cid:88) C(S, S, nτref ) ≈ andforthespecialcaseof τ syn (cid:28) τref (cid:1) ¯pkδ(cid:0)[n − k]τref (53) ek ln ¯pδ(cid:0)[n − k]τref (cid:1) . (54) Hence,sinceincreasingthemeanrate(orbias)oftheneu- ronleadstoanincreasein ¯pandthustoareduceddecay constant ln ¯p,moresignificantside-peaksemerge. For τ syn ≈ τref, the effective membrane distribution is not yet stationary and therefore, this approximation does not hold. To arrive at the exact solution, one would have torepeattheabovecalculationforallpossiblespiketime combinations,leadingtoarecursiveintegral[26].Further- more, one would also need to take into account the sit- uation where the first spike is itself part of a burst, i.e., is not the first spike in the burst. To circumvent a more tedious calculation, we use an approximation which is in between the two cases τ syn (cid:28) τref and τ syn ≈ τref: we use ¯p =(cid:82) ∞ ϑ duf (u, τref ϑ),whichprovidesareasonable approximationforshortbursts. 6.5 Cross-correlationoffreemembranepoten- tialsreceivingcorrelatedinput Similarly to the ACF of the membrane potential, one can calculatethecrosscorrelationfunctionoftheFMPsoftwo neurons receiving correlated noise input. First, the mem- branepotentialsaregivenby (55) (56) Thecovariancefunctioncanbewrittenas 0 + Stot,1 ∗ κ , 0 + Stot,2 ∗ κ . u1 = u1 u2 = u2 (cid:104)¯u1(t)¯u2(t + ∆)(cid:105) = (cid:104)u1(t)u2(t + ∆)(cid:105) − (cid:104)u1(t)(cid:105)(cid:104)u2(t)(cid:105) = (cid:104) − (cid:104) = ... (cid:0)Stot,1 ∗ κ(cid:1)(t)(cid:0)Stot,2 ∗ κ(cid:1)(t + ∆)(cid:105) (cid:0)Stot,1 ∗ κ(cid:1)(t)(cid:105)(cid:104) (cid:0)Stot,2 ∗ κ(cid:1)(t)(cid:105) =(cid:0) lim (cid:1)(∆) , (57a) (57b) (57c) with ¯u = u − (cid:104)u(cid:105), from which we obtain the crosscorre- lation function by normalizing with the product of stan- dard deviations of u1 and u2 (for notation, see Eq. (27)). The term containing the input correlation coefficient is (cid:104) ¯Stot,1(t) ¯Stot,2(t + ∆(cid:48))(cid:105). Plugging in the spike trains, we 1 T (cid:104) ¯Stot,1(t) ¯Stot,2(t + ∆(cid:48))(cid:105)T ∗ (cid:104)κ(t)κ(t + ∆(cid:48))(cid:105)∞ T→∞ 13 getfourcrosscorrelationterms (cid:88) · i(cid:105) i(cid:105) (cid:1) (cid:20) (∆) , (cid:104)gtot wlwm l,m∈{e,i} l,m∈{e,i} m − µ2 l − µ1 (cid:104)¯u1(t)¯u2(t + ∆)(cid:105) = ζ1ζ2 (cid:0)Erev (cid:1)(cid:0)Erev (cid:0)τ syn − (cid:104)τeff Λl,1Λm,2 (cid:104) ¯Sl,1(t) ¯Sm,2(t + ∆(cid:48))(cid:105) . (cid:104) ¯Stot,1(t) ¯Stot,2(t + ∆(cid:48))(cid:105) = (58) Sinceexcitatoryaswellasinhibitorynoiseinputsareran- domly drawn from the same pool of neurons, we can as- sumethat (cid:104) ¯Sl,1(t) ¯Sm,2(t + ∆(cid:48))(cid:105)isapproximatelyequalfor all combinations of synapse types when averaging over enough inputs, regardless of the underlying correlation structure/distributionofthenoisepool.Thefirstterm,how- ever, depends on the synapse types since the Λ-terms (Eq.(20b))containthedistancebetweenreversalpotentials andmeanFMP: (cid:88) (cid:104) ¯Sl,1(t) ¯Sm,2(t + ∆(cid:48))(cid:105) ∗ (cid:104)κ(t)κ(t + ∆(cid:48))(cid:105) (cid:104)Λe,1Λe,2(cid:105)inputs = −(cid:104)Λe,1Λi,2(cid:105)inputs , (cid:104)Λi,1Λi,2(cid:105)inputs = −(cid:104)Λi,1Λe,2(cid:105)inputs , (cid:21) (59) (cid:1). The cross- with constants ζi = τ syn correlationvanisheswhen,aftersummingovermanyinputs, thefollowingidentitieshold: (60a) (60b) where (cid:104)(·)(cid:105) is an average over all inputs, i.e., all neurons thatprovidenoise. Whilenotrelevantforoursimulations,itisworthnoting thattheexcitatoryandinhibitoryweightswithwhicheach neuroncontributesitsspiketrainscanberandomlydrawn fromnon-identicaldistributions.Byenforcingthefollowing correlationbetweenthenoiseweightsofbothneurons,one canintroduceaskewintotheweightdistributionwhichcom- pensatesforthedifferingdistancetothereversalpotentials: (61) A simple procedure to accomplish this is the follow- ing:First,wedrawtheabsoluteweights w1 and w2 from an arbitrary distribution and assign synapse types ran- domlywithprobabilities pe/iafterwards.If w2isexcitatory, we multiply w1 by , otherwise by .Thisway, (cid:104)w1(cid:105)remainsunchanged andtheresultingweightssufficeEq.(61). 6.6 Statespaceswitchfrom{0,1}to{-1,1} Toswitchfromthestatespace z ∈ {0, 1}to z(cid:48) ∈ {−1, 1} whileconservingthestateprobabilities(i.e., p(z) = p(z(cid:48))) rev − µ1)(Ee,2 (Ee,1 = −(Ee,1 e w2 rev − µ2)(cid:104)w1 rev − µ1)(Ei,2 rev − µ2)(cid:104)w1 e(cid:105)inputs e w2 rev−µ2+piEe,2 rev−µ2+piEe,2 i (cid:105)inputs peEi,2 peEi,2 rev−µ2 rev−µ2 rev−µ2 rev−µ2 Ee,2 Ei,2 14 i i j i i 2 4 i,j i,j i,j i,j i,j i,j = + = zi 1 2 1 2 1 2 − − 2 E(z(cid:48)) = (cid:88) (cid:88) onehastoadequatelytransformthedistributionparameters W and b. Since the distributions are of the form p(z) = exp(cid:0)zTWz + zTb(cid:1),thisisequivalenttorequiringthatthe energy E(z) = zTWz + zTbofeachstateremains,upto aconstant,unchanged. First,wecanwritetheenergyofastate z(cid:48)andusethe transformation z(cid:48) = 2z − 1toget (cid:88) (cid:88) (62a) z(cid:48)iW(cid:48)ijz(cid:48)j + z(cid:48)ib(cid:48)i (cid:18) (cid:88) (cid:88) ziW(cid:48)ijzj − 2 ziW(cid:48)ij − 2 W(cid:48)ijzj (cid:19) (cid:88) (cid:88) (cid:88) (62b) W(cid:48)ij b(cid:48)i + 2 zib(cid:48)i (cid:88) (cid:88) (cid:0)2b(cid:48)i zi4W(cid:48)ijzj + (cid:88) (cid:1) + C , (62c) W(cid:48)ij (cid:80) (cid:80) ib(cid:48)iandweused where Cisaconstant C = 1 i,jW(cid:48)ij − thefactthatW(cid:48)ij issymmetric.Sinceconstanttermsinthe energyleavetheprobabilitydistributioninvariant,wecan simplycompare E(z(cid:48))and E(z) (cid:88) (63) ibi , iWijzj + andextractthecorrectparametertransformation: (64) Wij = 4W(cid:48)ij , (65) bi = 2b(cid:48)i − 2 Fromthis,wecanalsocalculatetheinversetransformation rulefor z = 1 2 (z(cid:48) + 1): (66) W(cid:48)ij = (67) b(cid:48)i = 6.7 TranslationfromBoltzmanntoneurosynap- ticparameters Asdiscussedinthemethodssection,following[44],theacti- vationfunctionofLIFneuronsintheHCSisapproximately logisticandcanbewrittenas (68) where z/kisthestatevectorofallotherneuronsexceptthe k'th one and µ the mean membrane potential (Eq. (17b)). u0and αaretheinflectionpointandtheinverseslope,re- = (1 + exp (−(µ − u0)/α))−1 , p(zk = 1 z/k) = ϕ(µ) Wij , bi + W(cid:48)ij . Wij . (cid:88) (cid:88) E(z) = 1 4 1 2 zT zT 1 4 1 2 i,j j j i = j , 0 El = τm τeff Wkjzj − bk) −1 p(zk = 1 z/k) = (1 + exp (−bk))−1 , (cid:88) spectively.Furthermore,theconditionalprobability p(zk = 1 z/k) of a Boltzmann distribution over binary random variables zk,i.e., p(z) ∝ exp(cid:0) 1 2 zT W z + zT b(cid:1),isgivenby 1 + exp (− (cid:88) (69) withsymmetricweightmatrix W, Wii = 0 ∀i,andbiases b.Theseequationsallowatranslationfromtheparameters ofaBoltzmanndistribution (bi, Wij)toparametersofLIF neuronsandtheirsynapses (El, wij),suchthatthestate dynamicofthenetworkapproximatessamplingfromthetar- getBoltzmanndistribution. First,thebiases bcanbemappedtoleakpotentials El (or external currents) by requiring that, for W = 0 (that is, no synaptic input from other neurons), the activity of eachneuronequalstheconditionalprobabilityofthetarget Boltzmanndistribution (70) leadingtothetranslationrule (1 + exp (−(µ − u0)/α))−1 ! (71) TomapBoltzmannweights Wij tosynapticweights wij, we first have to rescale the Wij, as done for the biases in Eq. (71). However, leaky integrator neurons have non- rectangularPSPs,sotheirinteractionstrengthismodulated overtime.ThisisdifferentfromtheinteractioninBoltzmann machines,wherethePSPshapeisrectangular(Glauberdy- namics).Nevertheless,wecanderiveaheuristictranslation rulebyrequiringthatthemeaninteractionduringthere- fractory period of the presynaptic neuron is the same in bothcases,i.e.,(cid:90) τref (72a) (72b) where P SP (t)isgivenbyEq.(19).Fromthis,wegetthe translationruleforsynapticweights: (cid:16) (cid:17)(cid:16) (73) (cid:17) (cid:16) 1 − τ syn 7 Acknowledgments − τeff We thank Luziwei Leng, Nico Gurtler and Johannes Bill forvaluablediscussions.WefurtherthankEricMullerand Christian Mauch for maintenance of the computing clus- ter we used for simulations and Luziwei Leng for provid- (cid:17)−1 (cid:17)(cid:105) . Erev kj − µ e− τeff − 1 τref (αb + u0) − (cid:90) τref αWkjCm τref τ syn (cid:104)gsyn x (cid:105) gl Erev x . τ syn e− τref τ syn − 1 dt P SP (t) ! = dt αWij 0 = αWijτref , x∈{e,i} (cid:104) (cid:16) wkj = τeff ingcodeimplementingtheCASTalgorithm.Thisworkhas receivedfundingfromtheEuropeanUnion7thFramework Programmeundergrantagreement604102(HBP),theHori- zon 2020 Framework Programme under grant agreement 720270,785907(HBP)andtheManfredStarkFoundation. 8 References [1] GHHenry,POBishop,RMTupper,andBDreher.Ori- entationspecificityandresponsevariabilityofcellsin the striate cortex. Vision research, 13(9):1771 -- 1779, 1973. [2] PeterHSchiller,BarbaraLFinlay,andSusanFVol- man.Short-termresponsevariabilityofmonkeystriate neurons. Brainresearch,105(2):347 -- 349,1976. [3] Rufin Vogels, Werner Spileers, and Guy A Orban. Theresponsevariabilityofstriatecorticalneuronsin the behaving monkey. Experimental brain research, 77(2):432 -- 436,1989. [4] Robert J Snowden, Stefan Treue, and Richard A An- dersen. Theresponseofneuronsinareasv1andmtof thealertrhesusmonkeytomovingrandomdotpatterns. ExperimentalBrainResearch,88(2):389 -- 400,1992. [5] AmosArieli,AlexanderSterkin,AmiramGrinvald,and ADAertsen.Dynamicsofongoingactivity:explanation of the large variability in evoked cortical responses. Science,273(5283):1868,1996. [6] Rony Azouz and Charles M Gray. Cellular mecha- nisms contributing to response variability of cortical neuronsinvivo. JournalofNeuroscience,19(6):2209 -- 2223,1999. [7] Rajesh PN Rao, Bruno A Olshausen, and Michael S Lewicki. Probabilisticmodelsofthebrain:Perception andneuralfunction. MITpress,2002. [8] KonradPKordingandDanielMWolpert. Bayesian integration in sensorimotor Nature, learning. 427(6971):244 -- 247,2004. [9] Jan W Brascamp, Raymond Van Ee, Andre J Noest, RichardHAHJacobs,andAlbertVvandenBerg. The timecourseofbinocularrivalryrevealsafundamental roleofnoise. Journalofvision,6(11):8 -- 8,2006. [10] Gustavo Deco, Edmund T Rolls, and Ranulfo Romo. Stochastic dynamics as a principle of brain function. Progressinneurobiology,88(1):1 -- 16,2009. [11] J´ozsef Fiser, Pietro Berkes, Gergo Orb´an, and M´at´e Lengyel.Statisticallyoptimalperceptionandlearning: frombehaviortoneuralrepresentations.Trendsincog- nitivesciences,14(3):119 -- 130,2010. 15 [12] Wolfgang Maass. Searching for principles of brain computation. CurrentOpinioninBehavioralSciences, 11:81 -- 92,2016. [13] RichardBStein. Somemodelsofneuronalvariability. Biophysicaljournal,7(1):37 -- 68,1967. [14] NicolasBrunel. Dynamicsofsparselyconnectednet- works of excitatory and inhibitory spiking neurons. Journal of computational neuroscience, 8(3):183 -- 208, 2000. [15] NicolasFourcaudandNicolasBrunel.Dynamicsofthe firing probability of noisy integrate-and-fire neurons. Neuralcomputation,14(9):2057 -- 2110,2002. [16] Wulfram Gerstner, Werner M Kistler, Richard Naud, and Liam Paninski. Neuronal dynamics: From single neurons to networks and models of cognition. Cam- bridgeUniversityPress,2014. [17] DKSmettersandAnthonyZador. Synaptictransmis- sion:noisysynapsesandnoisyneurons. CurrentBiol- ogy,6(10):1217 -- 1218,1996. [18] Wolfgang Maass and Anthony M Zador. Dynamic stochastic synapses as computational units. In Ad- vancesinneuralinformationprocessingsystems,pages 194 -- 200,1998. [19] Peter N Steinmetz, Amit Manwani, Christof Koch, MichaelLondon,andIdanSegev.Subthresholdvoltage noise due to channel fluctuations in active neuronal membranes. Journal of computational neuroscience, 9(2):133 -- 148,2000. [20] YosefYaromandJornHounsgaard.Voltagefluctuations in neurons: signal or noise? Physiological reviews, 91(3):917 -- 929,2011. [21] Rub´en Moreno-Bote. Poisson-like spiking in circuits withprobabilisticsynapses. PLoScomputationalbiol- ogy,10(7):e1003522,2014. [22] Emre O Neftci, Bruno U Pedroni, Siddharth Joshi, Maruan Al-Shedivat, and Gert Cauwenberghs. Stochastic synapses enable efficient brain-inspired learningmachines. Frontiersinneuroscience,10:241, 2016. [23] CharlesFStevensandAnthonyMZador. Whenisan integrate-and-fireneuronlikeapoissonneuron?InAd- vancesinneuralinformationprocessingsystems,pages 103 -- 109,1996. [24] Hans E Plesser and Wulfram Gerstner. Noise in integrate-and-fireneurons:fromstochasticinputtoes- caperates. Neuralcomputation,12(2):367 -- 384,2000. [25] EJChichilnisky. Asimplewhitenoiseanalysisofneu- ronallightresponses.Network:ComputationinNeural Systems,12(2):199 -- 213,2001. [26] Wulfram Gerstner and Werner M Kistler. Spiking neuronmodels:Singleneurons,populations,plasticity. Cambridgeuniversitypress,2002. [27] Peter Dayan, LF Abbott, et al. Theoretical neuro- science:computationalandmathematicalmodelingof neural systems. Journal of Cognitive Neuroscience, 15(1):154 -- 155,2003. [28] SrdjanOstojicandNicolasBrunel. Fromspikingneu- ronmodelstolinear-nonlinearmodels.PLoScomputa- tionalbiology,7(1):e1001056,2011. [29] Gergo Orb´an, Pietro Berkes, J´ozsef Fiser, and M´at´e Lengyel. Neuralvariabilityandsampling-basedprob- abilisticrepresentationsinthevisualcortex. Neuron, 92(2):530 -- 543,2016. [30] LaurenceAitchisonandM´at´eLengyel.Thehamiltonian brain: efficient probabilistic inference with excitatory- inhibitoryneuralcircuitdynamics.PLoScomputational biology,12(12):e1005186,2016. [31] Moritz Deger, Moritz Helias, Clemens Boucsein, and Stefan Rotter. Statistical properties of superimposed stationaryspiketrains.JournalofComputationalNeu- roscience,32(3):443 -- 463,2012. [32] JINelson,PASalin,MH-JMunk,MArzi,andJBullier. Spatialandtemporalcoherenceincortico-corticalcon- nections:across-correlationstudyinareas17and18 inthecat. Visualneuroscience,9(1):21 -- 37,1992. [33] Bruno B Averbeck, Peter E Latham, and Alexandre Pouget. Neural correlations, population coding and computation. Nature reviews neuroscience, 7(5):358, 2006. [34] EmilioSalinasandTerrenceJSejnowski. Correlated neuronal activity and the flow of neural information. Naturereviewsneuroscience,2(8):539,2001. [35] RonenSegev,MorrisBenveniste,EyalHulata,Netta Cohen,AlexanderPalevski,EliKapon,YoashShapira, andEshelBen-Jacob.Longtermbehavioroflithograph- ically prepared in vitro neuronal networks. Physical reviewletters,88(11):118102,2002. [36] J´ozsefFiser,ChiayuChiu,andMichaelWeliky. Small modulationofongoingcorticaldynamicsbysensoryin- putduringnaturalvision.Nature,431(7008):573,2004. [37] RobertRosenbaum,TatjanaTchumatchenko,andRub´en Moreno-Bote.Correlatedneuronalactivityanditsrela- tionshiptocoding,dynamicsandnetworkarchitecture. Frontiersincomputationalneuroscience,8:102,2014. 16 Theory of [38] Rub´en Moreno-Bote, Alfonso Renart, and N´estor Parga. input spike auto-and cross- correlationsandtheireffectontheresponseofspiking neurons. Neuralcomputation,20(7):1651 -- 1705,2008. [39] Zachary F Mainen and Terrence J Sejnowski. Relia- bilityofspiketiminginneocorticalneurons. Science, 268(5216):1503,1995. [40] AnthonyZador. Impactofsynapticunreliabilityonthe informationtransmittedbyspikingneurons. Journalof Neurophysiology,79(3):1219 -- 1229,1998. [41] Alexander Rauch, Giancarlo La Camera, Hans-Rudolf Luscher,WalterSenn,andStefanoFusi. Neocortical pyramidalcellsrespondasintegrate-and-fireneurons toinvivo -- likeinputcurrents. Journalofneurophysiol- ogy,90(3):1598 -- 1612,2003. [42] Henry Markram, Eilif Muller, Srikanth Ramaswamy, MichaelWReimann,MarwanAbdellah,CarlosAguado Sanchez,AnastasiaAilamaki,LidiaAlonso-Nanclares, Nicolas Antille, Selim Arsever, et al. Reconstruc- tionandsimulationofneocorticalmicrocircuitry. Cell, 163(2):456 -- 492,2015. [43] Lars Buesing, Johannes Bill, Bernhard Nessler, and Wolfgang Maass. Neural dynamics as sampling: a modelforstochasticcomputationinrecurrentnetworks ofspikingneurons.PLoSComputBiol,7(11):e1002211, 2011. [44] Mihai A. Petrovici, Johannes Bill, Ilja Bytschok, Jo- hannes Schemmel, and Karlheinz Meier. Stochastic inferencewithspikingneuronsinthehigh-conductance state. Phys.Rev.E,94:042312,Oct2016. [45] Dejan Pecevski, Lars Buesing, and Wolfgang Maass. Probabilistic inference in general graphical models throughsamplinginstochasticnetworksofspikingneu- rons. PLoS computational biology, 7(12):e1002294, 2011. [46] Dimitri Probst, Mihai A Petrovici, Ilja Bytschok, Jo- hannesBill,DejanPecevski,JohannesSchemmel,and Karlheinz Meier. Probabilistic inference in discrete spacescanbeimplementedintonetworksoflifneurons. Frontiersincomputationalneuroscience,9,2015. [47] Luziwei Leng, Roman Martel, Oliver Breitwieser, Ilja Bytschok,WalterSenn,JohannesSchemmel,Karlheinz Meier, and Mihai A Petrovici. Spiking neurons with short-termsynapticplasticityformsuperiorgenerative networks. ScientificReports,8(1):10651,2018. [48] MichaelMonteforteandFredWolf.Dynamicfluxtubes formreservoirsofstabilityinneuronalcircuits. Physi- calReviewX,2(4):041007,2012. Spatiotempo- [49] Ryan Pyle and Robert Rosenbaum. ral dynamics and reliable computations in recurrent spiking neural networks. Physical review letters, 118(1):018103,2017. Idan Segev, Henry Markram, and [50] Galit Fuhrmann, Misha Tsodyks. Coding of temporal information by activity-dependent synapses. Journal of neurophysi- ology,87(1):140 -- 148,2002. [51] AndrewDavison,DanielBrÃČÂijderle,JochenEppler, Jens Kremkow, Eilif Muller, Dejan Pecevski, Laurent Perrinet,andPierreYger. Pynn:acommoninterface forneuronalnetworksimulators. FrontiersinNeuroin- formatics,2:11,2009. [52] Marc-Oliver Gewaltig and Markus Diesmann. Nest (neuralsimulationtool).Scholarpedia,2(4):1430,2007. [53] AlainDestexhe,MichaelRudolph,andDenisPar´e.The high-conductancestateofneocorticalneuronsinvivo. Naturereviewsneuroscience,4(9):739,2003. [54] ArvindKumar,SvenSchrader,AdAertsen,andStefan Rotter.Thehigh-conductancestateofcorticalnetworks. Neuralcomputation,20(1):1 -- 43,2008. [55] Solomon Kullback and Richard A Leibler. On infor- mation and sufficiency. The annals of mathematical statistics,22(1):79 -- 86,1951. [56] DavidHAckley,GeoffreyEHinton,andTerrenceJSe- jnowski. Alearningalgorithmforboltzmannmachines. Cognitivescience,9(1):147 -- 169,1985. [57] Ruslan Salakhutdinov. Learning deep boltzmann ma- chinesusingadaptivemcmc.InProceedingsofthe27th InternationalConferenceonMachineLearning(ICML- 10),pages943 -- 950,2010. [58] Johannes Schemmel, Johannes Fieres, and Karlheinz Meier. Wafer-scale integration of analog neural net- works. InNeuralNetworks,2008.IJCNN2008.(IEEE WorldCongressonComputationalIntelligence).IEEE International Joint Conference on, pages 431 -- 438. IEEE,2008. [59] Mihai A Petrovici, Bernhard Vogginger, Paul Muller, Oliver Breitwieser, Mikael Lundqvist, Lyle Muller, Matthias Ehrlich, Alain Destexhe, Anders Lansner, Ren´eSchuffny,etal. Characterizationandcompensa- tionofnetwork-levelanomaliesinmixed-signalneuro- morphicmodelingplatforms. PloSone,9(10):e108590, 2014. [60] Sebastian Schmitt, Johann Klahn, Guillaume Bellec, Andreas Grubl, Maurice Guettler, Andreas Hartel, StephanHartmann,DanHusmann,KaiHusmann,Se- bastian Jeltsch, et al. Neuromorphic hardware in the loop: Training a deep spiking network on the 17 brainscales wafer-scale system. In Neural Networks (IJCNN),2017InternationalJointConferenceon,pages 2227 -- 2234.IEEE,2017. [61] JakobJordan,MihaiAPetrovici,OliverBreitwieser,Jo- hannesSchemmel,KarlheinzMeier,MarkusDiesmann, andTomTetzlaff. Stochasticneuralcomputationwith- outnoise. arXivpreprintarXiv:1710.04931,2017. [62] Mihai A Petrovici, Sebastian Schmitt, Johann Klahn, DavidStockel,AnnaSchroeder,GuillaumeBellec,Jo- hannesBill,OliverBreitwieser,IljaBytschok,Andreas Grubl, et al. Pattern representation and recognition withacceleratedanalogneuromorphicsystems. InCir- cuits and Systems (ISCAS), 2017 IEEE International Symposiumon,pages1 -- 4.IEEE,2017. [63] Jun Haeng Lee, Tobi Delbruck, and Michael Pfeiffer. Trainingdeepspikingneuralnetworksusingbackprop- agation. Frontiersinneuroscience,10:508,2016. [64] FriedemannZenkeandSuryaGanguli.Superspike:Su- pervisedlearninginmultilayerspikingneuralnetworks. Neuralcomputation,30(6):1514 -- 1541,2018. [65] SaeedRezaKheradpisheh,MohammadGanjtabesh,Si- monJThorpe,andTimoth´eeMasquelier. Stdp-based spikingdeepconvolutionalneuralnetworksforobject recognition. NeuralNetworks,99:56 -- 67,2018. [66] Gregory Cohen, Saeed Afshar, Jonathan Tapson, and Andr´e van Schaik. Emnist: an extension of mnist to handwritten letters. arXiv preprint arXiv:1702.05373, 2017. [67] YannLeCun,L´eonBottou,YoshuaBengio,andPatrick Haffner. Gradient-basedlearningappliedtodocument recognition. Proceedings of the IEEE, 86(11):2278 -- 2324,1998. [68] A Aldo Faisal, Luc PJ Selen, and Daniel M Wolpert. Noise in the nervous system. Nature reviews neuro- science,9(4):292,2008. [69] Tiago Branco and Kevin Staras. The probability of neurotransmitterrelease:variabilityandfeedbackcon- trolatsinglesynapses.NatureReviewsNeuroscience, 10(5):373,2009. [70] John A White, Jay T Rubinstein, and Alan R Kay. Channel noise in neurons. Trends in neurosciences, 23(3):131 -- 137,2000. [71] H Sebastian Seung. Learning in spiking neural net- worksbyreinforcementofstochasticsynaptictransmis- sion. Neuron,40(6):1063 -- 1073,2003. [72] Xiaohui Xie and H Sebastian Seung. Learning in neuralnetworksbyreinforcementofirregularspiking. PhysicalReviewE,69(4):041909,2004. 18 [73] ZhangJChen,YongHe,PedroRosa-Neto,JurgenGer- mann,andAlanCEvans. Revealingmodulararchitec- tureofhumanbrainstructuralnetworksbyusingcor- ticalthicknessfrommri. Cerebralcortex,18(10):2374 -- 2381,2008. [74] EdBullmoreandOlafSporns.Complexbrainnetworks: graphtheoreticalanalysisofstructuralandfunctional systems. Nature Reviews Neuroscience, 10(3):186, 2009. [75] DavidMeunier,RenaudLambiotte,andEdwardTBull- more.Modularandhierarchicallymodularorganization of brain networks. Frontiers in neuroscience, 4:200, 2010. [76] Maxwell A Bertolero, BT Thomas Yeo, and Mark DâĂŹEsposito.Themodularandintegrativefunctional architecture of the human brain. Proceedings of the NationalAcademyofSciences,112(49):E6798 -- E6807, 2015. [77] SenSong,PerJesperSjostrom,MarkusReigl,Sacha Nelson, and Dmitri B Chklovskii. Highly nonrandom features of synaptic connectivity in local cortical cir- cuits. PLoSbiology,3(3):e68,2005. [78] WeiJiMa,JeffreyMBeck,PeterELatham,andAlexan- drePouget.Bayesianinferencewithprobabilisticpop- ulationcodes. Natureneuroscience,9(11):1432,2006. [79] Mark M Churchland, M Yu Byron, Stephen I Ryu, GopalSanthanam,andKrishnaVShenoy.Neuralvari- abilityinpremotorcortexprovidesasignatureofmo- torpreparation.JournalofNeuroscience,26(14):3697 -- 3712,2006. [80] MihaiAlexandruPetrovici.FormVersusFunction:The- ory and Models for Neuronal Substrates. Springer, 2016. [81] PeterHaunggiandPeterJung.Colorednoiseindynam- ical systems. Advances in chemical physics, 89:239 -- 326,1994. [82] ManuelOC´aceres.Harmonicpotentialdrivenbylong- rangecorrelatednoise.PhysicalReviewE,60(5):5208, 1999. [83] NorbertWiener. Generalizedharmonicanalysis. Acta mathematica,55(1):117 -- 258,1930. [84] Alexander Khintchine. Korrelationstheorie der sta- tionarenstochastischenprozesse. MathematischeAn- nalen,109(1):604 -- 615,1934. [85] IljaBytschok,DominikDold,JohannesSchemmel,Karl- heinzMeier,andMihaiA.Petrovici.Spike-basedprob- abilisticinferencewithcorrelatednoise. BMCNeuro- science,2017. wpre [86] IljaBytschok,DominikDold,JohannesSchemmel,Karl- heinzMeier,andMihaiAPetrovici.Spike-basedprob- abilisticinferencewithcorrelatednoise.arXivpreprint arXiv:1707.01746,2017. [87] LaurensvanderMaatenandGeoffreyHinton. Visual- izing data using t-sne. Journal of Machine Learning Research,9(Nov):2579 -- 2605,2008. 9 Supportinginformation 9.1 Experimentdetails AllsimulationsweredonewithPyNN0.8andNEST2.4.2. The LIF model was integrated with a time step of dt = 0.1 ms. Since the SSN is a time-continuous system, we could, in principle, retrieve a sample at every integration step.However,asindividualneuronsonlychangetheirstate onthetimescaleofrefractoryperiods,andhencenewstates emergeonasimilartimescale,wereadoutthestatesin intervalsof τref 2 .Ifnotstatedotherwise,weused USE = 1.0 and τrec = 10 ms as short term plasticity parameters for connectionswithineachSSNtoensurepostsynapticpoten- tialswithequalheight,asdiscussedin[80].Forbackground connections,i.e.,Poissoninputorbackgroundinputcoming fromotherSSNsinanensemble,weusedstaticsynapses (USE = 1.0and τrec → 0)insteadtofacilitatethemathe- maticalanalysis.Fortheinterconnectionsofanensemble, weexpectthatshort-termdepressionwillnotaltertheper- formanceofindividualSSNsinadrasticway,astheeffect willberathersmallonaverageifmostneuronsarefaraway fromtonicbursting.Thus,toallowaclearcomparabilityto SSNs receiving Poisson input, we chose to neglect short- termdepressionforensembleinterconnections.Theneuron parameters used throughout all simulations are listed in TableS1. 9.1.1 DetailstoFigure2B,C,Dofmaintext Theparametersforthetargetdistributionsofallnetworks were randomly drawn from beta distributions, i.e., W ∼ 2 · (beta(0.5, 0.5) − 0.5)and b ∼ 1.2 · (beta(0.5, 0.5) − 0.5). Thebiasofeachbackground-providingneuronwasadjusted toyieldthedesiredfiringrate p ∈ {0.1, 0.6, 0.9}incaseof (cid:17).Forthediffer- nosynapticinput(W = 0) b = log ent activity cases, we used N0.1 = 260, N0.6 = 50 and N0.9 = 34networksasbackgroundinputforeachneuron toreachthedesiredbackgroundnoisefrequency.ThePois- son frequency of the noise-providing networks was set to 3000 Hz.Activationfunctionswererecordedbyprovidingev- eryneuronwithbackgroundnoisefor 5 × 105 msandvary- ingitsleakpotential.Fortheautocorrelationfunctions,we first merged all individual noise spike trains and binned the resulting spike train with a bin size of 0.5 ms before calculatingtheautocorrelationfunction. (cid:16) p 1−p 19 ij ∼ 4 · (beta(0.5, 5.0) − 0.5) 9.1.2 DetailstoFigure2F,Gofmaintext Background input was generated from a pool of pairwise connected neurons (i.e., small subnetworks with two neu- ronseach)withstrongpositiveornegativeweightstoyield highlypositivelyornegativelycorrelatedspiketrains.Each pair of neurons in the main network (i.e, the network re- ceiving no Poisson input) received the spikes of 80 such subnetworksasbackgroundinput.Theweightsofthenoise- generatingsubnetworksweredrawnfrombetadistributions ij ∼ 4·(beta(5.0, 0.5)−0.5)(distributionstronglyskewed to positive weights) or wpre (skewedtonegativeweights).Theparametersofthemain network were randomly generated as in Fig. 2D. The ab- solute values of the weights Wnoise projecting from the noise-generating subnetworks to the main network were randomly generated from a (Gaussian-like) beta distribu- tion Wnoise ∼ (beta(4.0, 4.0) − 0.5) · 2 · 0.001 + 0.001 µS. The synapse type of each weight Wnoise was determined randomly with equal probability. Furthermore, inhibitory synapseswerescaledbyafactorof 1.35suchthatinhibitory andexcitatoryweightshavethesamemeanvalueasforthe simulationswithPoissonnoise(seeTableS1).Forthethree traces shown, the absolute value of each synaptic weight wasdrawnindependently.Synapsetypeswereeitherdrawn accordingtoapattern(Fig.2F,leftandmiddle)orindepen- dently(Fig.2F,right). 9.1.3 DetailstoFigure3A,B,CofmaintextandFigure Foreverynetwork,thetargetdistributionparameterswere againdrawnfromabetadistributionasinFig.2.Thecon- nectivityofthenoiseconnectionswassetto  = 0.05,i.e., eachneuronreceivedthespikesof 5%oftheremainingsub- networks' neurons as stochastic background input, for the ensembleof4003-neuronnetworks(notraining,FigureS2 intheSupportinginformation)andto  = 0.10fortheen- sembleof1006-neuronnetworks(training,maintext). The training was done for subnetworks with 6 neurons each, where every subnetwork was initialized with differ- ent weights and biases than the target parameters, also generated randomly. As an initial guess for the neurons' activation functions, we used the activation function of a neuronreceiving 2000 HzexcitatoryandinhibitoryPoisson input,leadingtoaslopeof α = 1.47 mVandamid-point at −52.97 mV.Theseparametersweresubsequentlyused totranslatetheweightandbiasupdatesgivenbytheHeb- bianwake-sleepalgorithm(seemaintext,Eqn.8and9)to updatesofsynapticweightsandleakpotentials.Thesub- networks were all trained simultaneously with contrastive divergence,wherethemodeltermwasapproximatedbysam- plingfor 1 × 105 ms.Thetrainingwasdonefor2000steps andwithalearningrateof t+2000.Asareference,50sub- networksreceivingonlyPoissonnoise(2000 Hz)werealso trainedinthesamewayfor2000steps. S2intheSupportinginformation 400 20 Self-activation of the network can be observed when a largeenoughfractionofneuronshaveasuprathresholdrest potential,inourcasearound30%. 9.1.4 DetailstoFigure3D,Eofmaintext The emulated ensemble consists of 15 4-neuron networks whichwererandomlyinitializedontwoHICANNchips(HI- CANN367and376onWafer33oftheBrainScaleSsys- tem).Theanaloghardwareparameterswhichdeterminethe physical range of weights adjustable by the 4bit setting weresetto gmax = 500and gmax div = 1.Giventhecur- rent state of development of the BrainScaleS system and itssurroundingsoftware,welimitedtheexperimenttosmall ensemblesinordertoavoidpotentialcommunicationbottle- necks.BiaseswereimplementedbyassigningeverySSNneu- ronabiasneuron,withitsrestpotentialsetabovethreshold toforcecontinuousspiking.Whileamoreresource-efficient implementation of biases is possible, this implementation allowed an easier mapping of neuron - bias pairs on the neuromorphic hardware. Bias strengths can then be ad- justed by modifying the synaptic weights between SSN neurons and their allocated bias neurons. The networks weretrainedwiththecontrastivedivergencelearningrule tosamplefromtheirrespectivetargetdistributions.Biases wererandomlydrawnfromanormaldistributionwith µ = 0 and σ = 0.25.Theweightmatriceswererandomlydrawn from W ∝ 2 · (beta(0.5, 0.5) − 0.5)andsubsequentlysym- metrized by averaging with their respecetive transposes 0.5 · (W + W T). Sincetherefractoryperiodsofhardwareneuronsvary,we further measured the refractory period of every neuron in theensemble,whichwaslaterusedtocalculatethebinary neuron states from spike raster plots. Refractory periods were measured by setting biases to large enough values todriveneuronstotheirmaximalfiringrate.Afterrunning the experiment, the duration of the refractory period can be approximated by dividing the experiment time by the numberofmeasuredspikes. During the whole experiment, the ensemble did not re- ceive external Poisson noise. individual SSNs received spikes from 20% of the remaining ensemble as backgroundinput,withnoiseconnectionshavinghardware weights of ±4 with the sign chosen randomly with equal probability.Translationbetweentheoreticaland4bithard- wareparameterswasdonebyclippingthevaluesintothe range [−15, 15]androundingtointegervalues.Calculation ofweightandbiasupdateswasperformedonahostcom- puter.Thelearningratewassetto η = 1.0forallnetworks performingbetterthanthemedianandtwicethisvaluefor networksperformingworse.Foreverytrainingstep,theen- semblewasrecordedfor 1 × 105 msbiologicaltimebefore applyingaparameterupdate. FortheexperimentswithPoissonnoise,everyneuronre- Instead, 20 intheSupportinginformation ceived 300 HzexternalPoissonnoiseprovidedbythehost computer. Theneuronparametersusedforallhardwareexperiments arelistedinTableS2. 9.1.5 DetailstoFigure4ofmaintextandFigureS4-6 To reduce the training time, we pretrained classical re- stricted Boltzmann machines on their respective datasets, followed by direct translation to spiking network parame- ters. To obtain better generative models, we utilized the CASTalgorithm[57]whichcombinescontrastivedivergence withsimulatedtempering.Eachsubnetworkwastrainedfor 200000stepswithaminibatchsizeof100,alearningrate of t+2000, an inverse temperature range β ∈ [1., 0.6] with 20 equidistant intervals and an adaptive factor γt = 9 1+t. Statesbetweenthefastandslowchainwereexchangedev- ery50samples.Tocollectthebackgroundstatisticsofthese subnetworks,wefirstsimulatedallnetworkswithstochastic Poissoninput.ToimprovethePoisson-stimulatedreference networks'mixingproperties,weutilizedshort-termdepres- sion to allow an easier escape from local energy minima faster (USE = 0.01, τrec = 280 ms, global weight rescale δW = 0.014−1). For classification, the grayscale value of imagepixelswastranslatedtospikingprobabilitiesofthe visibleunits,whichcanbeadjustedbysettingthebiases (cid:19). Spike probabilities of 0 and 1 were (cid:18) as ln mappedtobiasesof-50and50.Furthermore,duringclas- sification,theconnectionsprojectingbackfromthehidden neurons to the visible neurons were silenced in order to preventthehiddenlayerfrominfluencingtheclampedinput. Forpatternrivalry,thenon-occludedpixelswerebinarized. Intotal,eachSSNreceivedbackgroundinputfrom20%of the other networks' hidden neurons. For the classification results,theexperimentwasrepeated10timesfordifferent randomseeds(leadingtodifferentconnectivitymatricesbe- tween the SSNs). For training and testing, we used 400 and200imagesperclass.InFig.4D,consecutiveimages are 400 msapart.InFig.4E,fortheclamped"B",consecu- tiveimagesare2sapart,forthe"L"1.5s. The experiments with MNIST used an ensemble of five networkswith784visibleneuronsand500hiddenneurons each(Fig.S5andVideoS2intheSupportinginformation), trained on 6 · 103 images per digit class (where we took the digits provided by the EMNIST set to have balanced classes). Since the generative properties of larger SSNs depend heavily on the synaptic interaction, we also used short-term plasticity for the case without Poisson noise (USE = 0.01, τrec = 280 ms, δW = 0.01−1)toallowfluent mixing between digit classes [47]. For MNIST, each SSN receivedbackgroundinputfrom30%oftheothernetworks' hidden neurons. Furthermore, the excitatory noise weight wassetto wnoise The network-generated images were obtained by aver- grey value 1−grey value = 0.0009 µS. e agingthefiringactivityofthevisibleneurons(forpattern rivalry ±90 ms,forclassificationanddreamingofEMNIST ±80 msandforMNIST ±140 ms).Thetimeintervalswere chosentoreducetheblurcausedbymixingwhenplotting averagedspikingactivity. 9.2 Neuronparameters Table 1: Neuron parameters used in simulations through- outthemaintext.Themembranetimeconstantwaschosen smallsuchthatsmallernoisefrequenciessufficetoreacha high-conductancestate,allowingustousesmallerensem- blesandhencereducesimulationtime. 0.1nF 1.0ms -65.0mV 0.0mV -90.0mV -52.0mV -53.0mV 10.0ms 10.0ms 10.0ms 0.001 µS -0.00135 µS membranecapacitance membranetimeconstant leakpotential exc.reversalpotential inh.reversalpotential thresholdpotential resetpotential exc.synaptictimeconstant inh.synaptictimeconstant refractorytime exc.Poissonweights inh.Poissonweights Cm τm El Erev Erev e i ϑ  τ syn e τ syn i τref wnoise wnoise e i Cm τm El Erev Erev e Table2:Neuronparametersusedfortheimplementationsin anartificialneuralsubstrate.Notethattheseareintended parametersandtherealizedonescanvaryfromneuronto neuron. biasneurons 60.0mV -30.0mV 5.0ms 5.0ms 1.5ms ensembleneurons 0.2nF 7ms -20.0mV 60.0mV -100.0mV -20.0mV -35.0mV 8.0ms 8.0ms 4.0ms 9.3 Videocaptions 9.3.1 VideoS1 Sampled distribution over time from an autonomous en- semble (no explicit noise) of 15 4-neuron networks on an artificial neural substrate (the BrainScaleS system). (top) Median DKL of the ensemble (red) and the individual networks (red, transparent) as a function of time after training. The median DKL pre-training is shown in black. i ϑ  τ syn e τ syn i τref (Bottom) Comparison between target distribution (blue) and sampled distribution (red) for all networks. Most networksareabletoapproximatetheirtargetdistribution well(e.g.,thenetworksatposition(1,1),(4,1)and(4,2)with (row,column))oratleastapproximatethegeneralshapeof thetargetdistribution(e.g.,(0,1)and(0,2)).Thenetworks at position (2,2) and (3,0) show strong deviations from their respective target distributions due to single neuron deficiencies. Because of the speed-up of the BrainScaleS system,itonlytakes100mstoemulate 106msofbiological time. 9.3.2 VideoS2 Video of the data shown in Fig. S5 of the Supporting in- formation. An ensemble of five hierarchical networks with 784-500-10 (visible-hidden-label) neurons trained on the MNIST handwritten dataset generating digits without ex- plicitnoisesourcesisshown(insimulation).Everynetwork in the ensemble receives the spiking activity from hidden neuronsoftheothernetworksasstochasticinputonly.(top) Activityofthehiddenlayerofeachnetwork.(middle)Class labelcurrentlypredictedbythelabelneurons.(Bottom)Ac- tivityofthevisiblelayer.Togenerategrayscaleimagesfrom spikes,weaveragedthespikingactivityofeachneuronover awindowofsize ±90ms. 9.3.3 VideoS3 Video of the data shown in Fig. S7 of the Supporting information. A single hierarchical network with 784-200 (visible-hidden)neuronsgeneratingsamplesoftheMNIST handwrittendigitsdatasetwithoutexplicitnoisesourcesis shown(insimulation).Toinitializethenetwork,wetrained aBoltzmannmachineandtranslatedtheweightsandbiases toneurosynapticparameterstoreducesimulationtime.(top) Illustrationoftheusednetworkarchitecture.Lateral(non- plastic)connectionsineachlayerwereutilizedasanoise source(red),withaninterconnectivityof  = 0.2.(bottom) Averaged activity (average window ±90ms) of the visible layer(left)afterinitializingthenetwork,(middle)afterfur- thertrainingthenetworkand(right)forthecaseofexplicit Poissonnoiseinsteadoflateralinterconnections.Afterini- tialization,thenetworkisabletogeneraterecognizableim- agesbutdoesnotmixwellbetweendifferentdigitclasses. FurthertrainingthenetworkonimagesoftheMNISTtrain- ingsetimprovesbothimagequalityandmixing,rivalingthe quality of the reference setup with explicit Poisson noise. During the second training phase, neurosynaptic parame- tersareadjustedsuchthateveryneuronisabletoperform itstaskwiththeavailablebackgroundactivityitreceives. 21 9.4 Supportingfigures Figure S1: Compensation of input correlations by adjustment of weights and biases in an SSN [85, 86]. For simplicity, this is illustrated here for the caseofsharedinputcorrelations,buttheresultsholdforalltypesofstaticallycorrelatedinputs.(A)Exemplaryarchitectureofanetworkwith3neurons thatsamplesfromaBoltzmanndistributionwithparameters Wand b.Inordertoachievetherequiredstochasticregime,eachneuronreceivesexternal noiseintheformofPoissonspiketrains(notshown).(B)-(D)Exemplarysampleddistributionsforanetworkoftwoneurons.The"default"caseisthe onewhereallweightsandbiasesaresettozero(uniformdistribution,bluebars).(B)Sharednoisesourceshaveacorrelatingeffect,shiftingprobability massintothe(1,1)and(0,0)states(redbars).(C)Inthe {0, 1}2 space,increasedweightsintroducea(positive)shiftofprobabilitymassfromallother states towards the (1,1) state (red bars), which is markedly different from the effect of correlated noise. (D) In the {−1, 1}2 space, increased weights have the same effect as correlated noise (red bars). (E) Dependence of the correlation coefficient r between the states of two neurons on the change in synaptic weight ∆W (cid:48) (red) and the shared noise ratio s (blue). These define bijective functions g and h that can be used to compute the weight change (∆W (cid:48) = f (s), with f := g−1 ◦ h) needed to compensate the effect of correlated noise in the {−1, 1}N space. (F) Study of the optimal compensationruleinanetworkwithtwoneurons.Forsimplicity,theordinaterepresentsweightchangesforanetworkwithstatesinthe {−1, 1}2space, whicharethentranslatedtocorrespondingparameters(W, b)forthe {0, 1}2 statespace.Thecolormapshowsthedifferencebetweenthesampledand (cid:0)pnet (cid:107) ptarget(cid:1).Themappingprovidedbythecompensationrule f (see(E)) thetargetdistributionmeasuredbytheKullback-Leiblerdivergence DKL isdepictedbythegreencurve.Notethatthecompensationrule ∆W (cid:48) = f (s)providesanearlyoptimalparametertranslation.Remainingdeviationsare duetodifferencesbetweenLIFandGlauberdynamics.(G)CompensationofnoisecorrelationsinanSSNwithtenneurons.Theresultsaredepictedfor asetoftenrandomlydrawnBoltzmanndistributionsover z ∈ {0, 1}10 (errorbars).ForasetofrandomlychosenBoltzmanndistributions,aten-neuron networkperformssamplinginthepresenceofpairwise-sharednoiseratios s(x-axis).Thebluelinemarksthesamplingperformancewithoutnoise-induced correlations(s = 0).Foranincreasingsharednoiseratio,uncompensatednoise(green)inducesasignificantincreaseinsamplingerror.Aftercompensation, thesamplingperformanceisnearlycompletelyrestored.Asbefore,remainingdeviationsareduetodifferencesbetweenLIFandGlauberdynamics.(H)An LIF-basedten-neuronnetworkwithsharednoisesources(s = 0.3foreachneuronpair)istrainedwithdatasamplesgeneratedfromatargetBoltzmann distribution(bluebars).Duringtraining,thesampleddistributionbecomesanincreasinglybetterapproximationofthetargetdistribution(redline).For comparison,wealsoshowthedistributionsampledbyanSSNwithparameterstranslateddirectlyfromtheBoltzmannparameters(purple).Thetrained networkisabletoimproveuponthisresultbecauselearningimplicitlycompensatesfortheabovementioneddifferencesbetweenLIFandGlauberdynamics. 22 z1z2z3W=WTb1b2b3H0.00.10.20.30.40.5shared noise ratio s00.20.40.60.8corr. coeff. rhg-1E00011011z10-1p(z)00011011z10-1p(z)-1-1-111-111z010-1p(z0)10-210-1p(z)0500100015002000training steps10-210-1100DKL(pnetpdata)0.00.10.20.30.40.5shared noise ratio s10-210-1GA0.00.10.20.30.4shared noise ratio s-0.36-0.24-0.120.00.12∆W0FDBC0.00.10.20.30.40.50.6∆W010-310-210-1DKL(pnetptarget)z1z2zk∈{0,1}W12=W21>0z1z2z0k∈{−1,1}W012=W021>0z01z02 FigureS2:Translatingtheparametersoftargetdistributionstoneurosynapticparameters.(A)Astraightforwardwaytosetuptheparametersofeach network(wij and El)istousetheparametertranslationasdescribedinthemaintext,i.e.,usethecorrespondingactivationfunctionofeachneuronto correctlyaccountforthebackgroundnoisestatistics.Thisisdemonstratedhereforthecaseof(left)399networks(onlytwoshown)receivingPoisson noiseandonenetworkonlyreceivingensembleinputand(right)allnetworksonlyreceivingensembleinput.Inbothcases,theresultingactivationfunction isthesameandwecanindeeduseittotranslatetheparametersofthetargetdistributiontoneurosynapticparameters.(B)Usingthecorresponding activationfunctionstosetuptheensemble(butnotraining),eachnetworkintheensembleisindeedabletoaccuratelysamplefromitstargetdistribution withoutexplicitnoise,asexpectedfromourconsiderationsin(A)andthemaintext.Thisisshownhere(insoftwaresimulations)foranensembleof400 3-neuronSSNswithaninterconnectionprobabilityof 0.05,reachingamedian DKL of 12.8+6.4−5.0 × 10−3 (blue),whichisclosetotheidealresultwith Poissonnoiseof 6.2+2.0−2.0 × 10−3(black,errorsgivenasthefirstandthirdquartile). FigureS3:TheBrainScalesneuromorphicsystem.(A)AsingleHICANNchip(HighInputCountAnalogNeuralNetwork),theelementalbuildingblock of the BrainScaleS wafer. The HICANN consists of two symmetric halves and harbors analog implementations of adaptive exponential integrate-and- fire(AdEx)neuronsandconductance-basedsynapsesin180nmCMOStechnology.Floatinggatesnexttotheneuroncircuitsareusedtostoreneuron parameters.Spikesarerouteddigitallythroughhorizontalandverticalbuses(notshown)andtranslatedintopostsynapticconductancesinthesynapse array. Unlike in simulations on general-purpose CPUs, here neurons and synapses are physically implemented, with no numeric computations being performedtocalculatenetworkdynamics.Asinglewaferconsistsof384HICANNchips.(B)IndividualcomponentsoftheBrainScaleSsystem,including bothwaferandsupportstructure.Forinstance,FPGAboardsprovideanI/OinterfaceforwaferconfigurationandspikedataandGiga-Ethernetslots provideaconnectionbetweenFPGAsandthecontrolclusterfromwhichusersconducttheirexperimentsviaPythonscriptsusingthePyNNAPI.(C) CompletelyassembledwaferoftheBrainScaleSneuromorphicsystem. 23 A101102103104105106sim. time [ms]103102101100DKL(pnetptarget)EnsemblePoisson referenceBBCA Figure S4: t-SNE representation [87] of consecutively generated images of two of the four SSNs trained on EMNIST. Both SSNs smoothly traverse severalregionsofthestatespacerepresentingimageclasseswhiledreaming.Thereddiamondmarksthefirstimageinthesequence,graylinesconnect consecutiveimages.Consecutiveimagesare 200 msapart. 24 FigureS5:DreamingofMNISTinanensembleofnetworks.(A)DreamingensembleoffivehierarchicalSSNswith784visible,500hiddenand10label neurons(withoutexplicitnoise).Eachrowrepresentssamplesfromasinglenetworkoftheensembles,withsamplesbeing 375 msapart.Tosetupthe ensemble,arestrictedBoltzmannmachinewastrainedontheMNISTdatasetandtheresultingparameterstranslatedtocorrespondingneurosynaptic parametersoftheensemble.Here,tofacilitatemixing,weusedshort-termdepressiontomodulatesynapticinteractionsandweakenattractorstatesthat wouldbeotherwisedifficulttoescape[47].(B)t-SNErepresentation[87]ofconsecutivelygeneratedimagesoftwoofthefiveSSNstrainedonMNIST digits.BothSSNsareabletogenerateandmixbetweendiverseimagesofdifferentdigitclasseswhiledreaming.Thereddiamondmarksthefirstimage inthesequence,graylinesconnectconsecutiveimages.Consecutiveimagesare 400 msapart. 25 BA FigureS6:PatterncompletioninanensembletrainedonEMNIST.(A)Relativeabundanceofthelabeloutputwhileclampingpartsofa"B".Mostofthe time(79.85%),theimageiscorrectlyclassifiedasa"B".Theclosestalternativeexplanation,an"R",isgeneratedsecondmost(17.45%).Theremaining classesareexploredsignificantlylessoftenbythenetwork(0.43%,0.70%,1.57%).(B)Examplesofthevisiblelayeractivitywhilethelabellayerclassifies thepartiallyclampedimageseitherasa"B"(top)oran"R"(bottom).(C)Examplesofthevisiblelayeractivitywhileclassifyingtheimageasa"T","X" or"V".Inthesecases,theimagesgeneratedbythevisibleneuronsshowprominentfeaturesoftheseletters. 26 V102101100relative abundanceBRTXVlabel predictionABBCRTX Figure S7: A single hierarchical network with 784-200 (visible-hidden) neurons generating samples of the MNIST handwritten digits dataset without explicitnoisesources(insimulation),representedviat-SNE.(A)Illustrationoftheusednetworkarchitecture.Lateral(non-plastic)connectionsineach layerwereutilizedasanoisesource(red),withaninterconnectivityof  = 0.2.(B,C)Averagedactivity(averagewindow ±90ms)ofthevisiblelayer(B) afterinitializingthenetworkand(C)afterfurthertrainingthenetwork.Afterinitialization,thenetworkisabletogeneraterecognizableimagesbutdoes notmixwellbetweendifferentdigitclassessincethenetworkisnotabletocorrectlyutilizeitsownbackgroundactivityasnoiseyet(B).Furthertraining thenetworkonimagesoftheMNISTtrainingset(usingstandardContrastiveDivergencewithbatchsize 100,learningrate t+2000 with tthenumberof updates,for 1000trainingupdatesandapresentationtimepertrainingsampleof 200ms)improvesbothimagequalityandmixing(C).Duringthesecond trainingphase,neurosynapticparametersareadjustedsuchthateveryneuronisabletoperformitstaskwiththeavailablebackgroundactivityitreceives. 40 27 CBA
1608.03616
1
1608
2016-08-11T20:48:18
Benchmarking confound regression strategies for the control of motion artifact in studies of functional connectivity
[ "q-bio.NC" ]
Since initial reports regarding the impact of motion artifact on measures of functional connectivity, there has been a proliferation of confound regression methods to limit its impact. However, recent techniques have not been systematically evaluated using consistent outcome measures. Here, we provide a systematic evaluation of 12 commonly used confound regression methods in 193 young adults. Specifically, we compare methods according to three benchmarks, including the residual relationship between motion and connectivity, distance-dependent effects of motion on connectivity, and additional degrees of freedom lost in confound regression. Our results delineate two clear trade-offs among methods. First, methods that include global signal regression minimize the relationship between connectivity and motion, but unmask distance-dependent artifact. In contrast, censoring methods mitigate both motion artifact and distance-dependence, but use additional degrees of freedom. Taken together, these results emphasize the heterogeneous efficacy of proposed methods, and suggest that different confound regression strategies may be appropriate in the context of specific scientific goals.
q-bio.NC
q-bio
Benchmarking confound regression strategies for the control of motion artifact in studies of functional connectivity Rastko Cirica, Daniel H. Wolfa, Jonathan D. Powerb, David R. Roalfa, Graham Bauma, Kosha Ruparela, Russell T. Shinoharac, Mark A. Elliottd, Simon B. Eickhoffe,f, Christos Davatzikosd, Ruben C. Gura, Raquel E. Gura, Danielle S. Bassettg,h, Theodore D. Satterthwaitea,∗ aDepartment of Psychiatry, Perelman School of Medicine, University of Pennsylvania, Philadelphia PA, USA bSackler Institute for Developmental Psychobiology, Weill Medical College of Cornell University, New York, NY, USA cDepartment of Biostatistics and Epidemiology, Perelman School of Medicine, University of Pennsylvania, Philadelphia, PA, USA dDepartment of Radiology, Perelman School of Medicine, University of Pennsylvania, Philadelphia PA, USA eDepartment of Clinical Neuroscience and Medical Psychology, Heinrich-Heine University Dusseldorf fInstitute of Neuroscience and Medicine (INM-1), Research Center Julich gDepartment of Bioengineering, University of Pennsylvania, Philadelphia PA, USA hDepartment of Electrical and Systems Engineering, University of Pennsylvania, Philadelphia PA, USA Abstract Since initial reports regarding the impact of motion artifact on measures of functional connectivity, there has been a proliferation of confound regression methods to limit its impact. However, recent techniques have not been systematically evaluated using consistent outcome measures. Here, we provide a systematic evaluation of 12 commonly used confound regression methods in 193 young adults. Specifically, we compare methods according to three benchmarks, including the residual relationship between motion and connectivity, distance-dependent effects of motion on connectivity, and additional degrees of freedom lost in confound regression. Our results delineate two clear trade-offs among methods. First, methods that include global signal regression minimize the relationship between connectivity and motion, but unmask distance-dependent artifact. In contrast, censoring methods mitigate both motion artifact and distance-dependence, but use additional degrees of freedom. Taken together, these results emphasize the heterogeneous efficacy of proposed methods, and suggest that different confound regression strategies may be appropriate in the context of specific scientific goals. Keywords: fMRI, functional connectivity, artifact, confound, motion, noise Introduction Resting-state (intrinsic) functional connectivity (rsfc- MRI) has evolved to become one of the most common brain imaging phenotypes (Craddock et al., 2013; Fox and Raichle, 2007; Power et al., 2014b; Smith et al., 2013; Van Dijk et al., 2010), and has been critical for understand- ing fundamental properties of brain organization (Damoi- seaux et al., 2006; Fox et al., 2005; Power et al., 2011; Yeo et al., 2011), brain development over the lifespan (Di- Martino et al., 2014; Dosenbach et al., 2011; Fair et al., 2008), and abnormalities associated with diverse clinical conditions (Baker et al., 2014; Buckner et al., 2008; Fair et al., 2010). rsfc-MRI has numerous advantages, includ- ing ease of acquisition and suitability for a wide and ex- panding array of analysis techniques. However, despite knowledge that in-scanner motion can influence measures of activation from task-related fMRI (Friston et al., 1996), ∗Corresponding author Email address: [email protected] (Theodore D. Satterthwaite ) Preprint submitted to arXiv the impact of in-scanner motion on measures of functional connectivity was not explored for 16 years after its initial discovery. However, since the near-simultaneous publica- tion of three independent reports in early 2012 (Van Dijk et al., 2012; Power et al., 2012; Satterthwaite et al., 2012), it has been increasingly recognized that motion can have a large impact on rsfc-MRI measurements, and can system- atically bias inference. This bias is particularly problem- atic in developmental or clinical populations where mo- tion is correlated with the independent variable of inter- est (age, diagnosis) (Satterthwaite et al., 2012; Fair et al., 2012), and has resulted in the re-evaluation of numerous published findings. In response to this challenge, there has been a recent proliferation of methods aimed at mitigating the impact of motion on functional connectivity (Yan et al., 2013a; Power et al., 2015). These methods can be broadly grouped into several categories. First, high-parameter con- found regression strategies use expansions of realignment parameters or tissue-compartment signals, often including derivative and quadratic regressors (Friston et al., 1996; Satterthwaite et al., 2013; Yan et al., 2013a). Second, prin- August 15, 2016 cipal component analysis (PCA) based methods (Comp- Cor; Behzadi et al. (2007); Muschelli et al. (2014)) find the primary directions of variation within high-noise areas de- fined by anatomy (e.g., aCompCor) or temporal variance (tCompCor). Third, whole-brain independent component analysis (ICA; Beckmann et al. (2005)) of single-subject time series has increasingly been used for de-noising, with noise components selected either by a trained classifier (ICA-FIX; Griffanti et al. (2014); Salimi-Khorshidi et al. (2014)) or using a priori heuristics (ICA-AROMA; Pruim et al. (2015b,a)). Fourth, temporal censoring techniques identify and remove (or de-weight) specific volumes con- taminated by motion artifact, often followed by interpo- lation. These techniques include scrubbing (Power et al., 2012, 2014a, 2015), spike regression (Satterthwaite et al., 2013), and de-spiking (Jo et al., 2013; Patel et al., 2014). Censoring techniques have been reported to attenuate mo- tion artifact, but at the cost of a shorter time series and variably reduced degrees of freedom. Fifth, one recent re- port emphasized the relative merits of spatially-tailored confound regression using local white matter signals (wm- Local; Jo et al. (2013)). Finally, the inclusion of global signal regression (GSR) (Macey et al., 2004) in confound regression models remains a source of controversy (Fox et al., 2009; Murphy et al., 2009; Chai et al., 2012; Saad et al., 2012; Yan et al., 2013c). While several studies have suggested its utility in de-noising (Fox et al., 2009; Power et al., 2015; Satterthwaite et al., 2013; Yan et al., 2013a), other studies have emphasized the risk of removing a valu- able signal (Yang et al., 2014; Hahamy et al., 2014), po- tentially biasing group differences (Gotts et al., 2013; Saad et al., 2012), or exacerbating distance-dependent motion artifact. Distance-dependent artifact (Power et al., 2012; Satterthwaite et al., 2012) manifests as increased connec- tivity in short-range connections, and reduced connectiv- ity in long-range connections, which has the potential to impact measures of network topology (Yan et al., 2013b). This recent proliferation of de-noising techniques has prompted excitement but also sowed confusion. Unsurpris- ingly, new de-noising pipelines have often tended to em- phasize outcome measures that suggest their relative supe- riority. As a result, investigators often anecdotally report substantial uncertainty regarding which pipeline should be used. Such uncertainty has been exacerbated by the lack of common outcome measures used across studies, which has hampered direct comparison among pipelines. While one review paper has summarized recent developments in this rapidly-evolving sub-field (Power et al., 2015), systematic evaluation of de-noising pipelines according to a range of benchmarks remains lacking. Accordingly, in this report we compare a dozen of the most commonly used confound regression strategies in a large (N = 193) dataset of young adults. Pipelines eval- uated include standard techniques, high-parameter con- found regression, PCA-based techniques such as aComp- Cor and tCompcor, ICA-based approaches such as ICA- AROMA, spatially-tailored local white matter regression, and three different censoring techniques (spike regression, de-spiking, and scrubbing); GSR is included in many pipelines as well. Critically, we compare these pipelines ac- cording to three intuitive benchmarks, including the resid- ual relationship between functional connectivity and sub- ject motion, the degree of distance-dependent artifact, and the loss of temporal degrees of freedom. As described be- low, results underscore the relative strengths and weak- nesses among these methods, and outline clear trade-offs among commonly used confound regression approaches. Materials and methods Participants and data acquisition The task-free BOLD data used in this study (N = 193) were selected from the Philadelphia Neurodevelopmental Cohort (PNC) (Satterthwaite et al., 2014, 2016) on the ba- sis of age, health, and data quality. In order to minimize the potential impact of developmental effects, subjects se- lected for inclusion were aged at least 18 years at the date of scan. All participants were free of medical problems that could impact brain function (Merikangas et al., 2010), lacked gross structural brain abnormalities (Gur et al., 2013), and were not taking psychotropic medication. Fur- thermore, subjects were excluded from the study if they failed to satisfy any of three criteria for functional image quality: if subject movement (relative root mean square displacement) averaged over all frames exceeded 0.2mm; if over 20 individual frames featured movement in excess of 0.25mm (Satterthwaite et al., 2012); or if brain cov- erage during acquisition was incomplete. As a result of these criteria, participants with gross in-scanner motion were excluded, allowing us to evaluate the utility of con- found regression strategies for the mitigation of artifact due to micro-movements. Structural and functional subject data were acquired on a 3T Siemens Tim Trio scanner with a 32-channel head coil (Erlangen, Germany), as previously described (Satterthwaite et al., 2014, 2016). High-resolution struc- tural images were acquired in order to facilitate align- ment of individual subject images into a common space. Structural images were acquired using a magnetization- prepared, rapid-acquisition gradient-echo (MPRAGE) T1- weighted sequence (TR = 1810ms; TE = 3.51ms; FoV = 180 × 240mm; resolution 1mm isotropic). Approxi- mately 6min of task-free functional data were acquired for each subject using a blood oxygen level-dependent (BOLD-weighted) sequence (TR = 3000ms; TE = 32ms; FoV = 192 × 192mm; resolution 3mm isotropic; 124 spa- tial volumes). Prior to scanning, in order to acclimate subjects to the MRI environment and to help subjects learn to remain still during the actual scanning session, a mock scanning session was conducted using a decom- missioned MRI scanner and head coil. Mock scanning was accompanied by acoustic recordings of the noise pro- duced by gradient coils for each scanning pulse sequence. 2 During these sessions, feedback regarding head movement was provided using the MoTrack motion tracking system (Psychology Software Tools, Inc, Sharpsburg, PA). Motion feedback was only given during the mock scanning session. In order to further minimize motion, prior to data acquisi- tion subjects' heads were stabilized in the head coil using one foam pad over each ear and a third over the top of the head. During the resting-state scan, a fixation cross was displayed as images were acquired. Subjects were in- structed to stay awake, keep their eyes open, fixate on the displayed crosshair, and remain still. Structural image processing A study-specific template was generated from a sam- ple of 120 PNC subjects balanced across sex, race, and age bins using the buildTemplateParallel procedure in ANTs (Avants et al., 2011a). Study-specific tissue priors were created using a multi-atlas segmentation procedure (Wang et al., 2014). Next, each subject's high-resolution structural image was processed using the ANTs Cortical Thickness Pipeline (Tustison et al., 2014). Following bias field correction (Tustison et al., 2010), each structural im- age was diffeomorphically registered to the study-specific PNC template using the top-performing SyN deformation (Klein et al., 2009). Study-specific tissue priors were used to guide brain extraction and segmentation of the subject's structural image (Avants et al., 2011b). BOLD time series processing Task-free functional images were processed using the XCP Engine (Ciric et al., In Preparation), which was con- figured to support the 12 pipelines evaluated in this study (see Figure 1). Each pipeline was based on de-noising strategies previously described in the neuroimaging lit- erature. A number of preprocessing procedures were in- cluded across all de-noising pipelines. Common elements of preprocessing included (1) correction for distortions in- duced by magnetic field inhomogeneities, (2) removal of the 4 initial volumes of each acquisition, (3) realignment of all volumes to a selected reference volume using mcflirt (Jenkinson et al., 2002), (4) demeaning and removal of any linear or quadratic trends, (5) co-registration of func- tional data to the high-resolution structural image using boundary-based registration (Greve and Fischl, 2009), and (6) temporal filtering using a first-order Butterworth fil- ter with a passband between 0.01 and 0.08 Hz. These preliminary processing stages were then followed by the confound regression procedures described below. In or- der to prevent frequency-dependent mismatch during con- found regression (Hallquist et al., 2013), all regressors were band-pass filtered to retain the same frequency range as the data. As in our prior work (Satterthwaite et al., 2012, 2013), the primary summary metric of subject motion used was the mean relative root-mean-square displacement cal- culated during time series realignment using mcflirt. Overview of confound regression strategies The primary objective of the current study was to evalu- ate the performance of common de-noising strategies. We selected 12 de-noising models, labelled 1–12 below, for evaluation (Figure 1). Models 1–5 used nuisance parame- ters derived from 6 movement estimates and 3 physiolog- ical time series, as well as their temporal derivatives and quadratic expansions. • Model 1. (2P) Used only the 2 physiological time se- ries: mean signal in WM and mean signal in CSF, and functioned as a base model for comparison to other more complex confound regression models. • Model 2. (6P) Used only the 6 motion estimates de- rived from mcflirt realignment as explanatory vari- ables. • Model 3. (9P) Combined the 6 motion estimates and 2 physiological time series with global signal re- gression. This model has been widely applied to func- tional connectivity studies (Fox et al., 2005, 2009). • Model 4. (24P) Expansion of model 2 that in- cludes 6 motion parameters, 6 temporal derivatives, 6 quadratic terms, and 6 quadratic expansions of the derivatives of motion estimates for a total 24 regres- sors (Friston et al., 1996). • Model 5. (36P) Similar expansion of model 3: 9 regressors, their derivatives, quadratic terms, and squares of derivatives (Satterthwaite et al., 2013). Models 6–8 further expanded upon this maximal 36P strategy by incorporating censoring approaches. • Model 6. (36P+despike) Included 36 regressors as well as despiking (Cox, 1996). • Model 7. (36P+spkreg) Included 36 regressors as well as spike regression, as in Satterthwaite et al. (2013). • Model 8. (36P+scrub) Included 36 regressors as well as motion scrubbing, as in Power et al. (2014a). Models 9 and 10 adapted variants of the PCA-based CompCor approach. • Model 9. (aCompCor) Used 5 principal components each from the WM and CSF, in addition to motion estimates and their temporal derivatives (Muschelli et al., 2014). • Model 10. (tCompCor) Used 6 principal compo- nents from high-variance voxels (Behzadi et al., 2007). The final two models evaluated used a regressor de- rived from local WM signals, and subject-specific ICA de- noising, respectively. 3 • Model 11. (wmLocal) Used a voxelwise, localised WM regressor in addition to motion estimates and their temporal derivatives and despiking (Jo et al., 2013). • Model 12. (ICA-AROMA) Used a recently devel- oped ICA-based procedure for removal of motion- related variance from BOLD data, together with mean WM and CSF regressors (Pruim et al., 2015a,b). We explicitly limited our scope to models that did not require training a classifier, and did not evaluate confound regression strategies that require extensive parameter op- timization (Salimi-Khorshidi et al., 2014; Griffanti et al., 2014; Patel et al., 2014). Furthermore, in order to con- strain the parameter space, we did not examine unpub- lished combinations of de-noising approaches. Confound regression using realignment parameters Time series of six realignment parameters (three transla- tional and three rotational) for each subject were returned by mcflirt as part of time series realignment (motion cor- rection). Additionally, the temporal derivative, quadratic terms, and quadratic of the temporal derivative of each of the realignment parameters were calculated, yielding 24 re- alignment regressors in total. The original six realignment parameters were included in confound regression models 2 and 3. Models 9 and 11 included 12 realignment regressors – the 6 realignment parameters and their temporal deriva- tives – while the full set of 24 expanded realignment regres- sors were included as part of confound regression models 4–8. Global signal regression The mean global signal was computed by averaging across all voxelwise time series located within a subject- specific mask covering the entire brain. The global signal was included in model 3, while the expanded models 5–8 included 4 global signal regressors: the global signal, its derivative, its square, and the derivative of its square. Tissue class regressors Mean white matter (WM) and cerebrospinal fluid (CSF) signals were computed by averaging within masks derived from the segmentation of each subject's structural image; these masks were eroded using AFNI's 3dmask tool (Cox, 1996) to prevent inclusion of gray matter signal via partial- volume effects. The WM mask was eroded at the 2-voxel level, while the CSF mask was eroded at the 1-voxel level. More liberal erosion often resulted in empty masks in our data. Temporal derivatives, quadratic terms, and squares of the derivative were computed as above. Two tissue class regressors (WM and CSF) were included in models 3 and 12, whereas their expansions (8 regressors) were included in models 4–8. Local white matter regression Model 11 used a local WM regressor (Jo et al., 2013). This was computed in AFNI using 3dLocalstat (Cox, 1996). Unlike the regressors described above, which were voxel-invariant, the value of the local WM regressor was computed separately at each voxel. For each voxel, a sphere of radius 45mm was first centered on that voxel; this sphere defined that voxel's local neighborhood. Next, this spherical neighborhood was intersected with an eroded WM mask to produce a local WM mask, which included only the fraction of the WM that was also in the voxel's neighborhood. The mean signal within this new local WM mask was then used to model the local WM signal at the voxel (Jo et al., 2013). This process was repeated at every voxel in order to generate the local WM regressor. This local WM regressor was included in model 11 along with realignment parameters and their derivatives (12 total); this model also included voxelwise de-spiking. CompCor Principal component analysis (PCA) can be used to model noise in BOLD time series (Behzadi et al., 2007; Muschelli et al., 2014). Broadly, the use of PCA-derived regressors to model noise is called component-based cor- rection (CompCor). Numerous variants of CompCor have been developed; here, our focus will be on anatomical CompCor (aCompCor, model 9) and temporal CompCor (tCompCor, model 10). In aCompCor, a PCA is per- formed within an anatomically defined tissue class of in- terest. We extracted 5 components for WM and CSF each, yielding 10 compcor components (Muschelli et al., 2014). As part of model 9, as in Muschelli et al. (2014), these 10 aCompCor components were combined with 12 re-alignment parameters (raw and temporal derivative). In tCompCor, the temporal variance of the BOLD signal is first computed at each voxel. Subsequently, a mask is generated from high-variance voxels, and principal compo- nents are extracted from the time series at these voxels. In confound regression model 10, tCompcor was imple- mented using ANTs, with 6 tCompCor components used as confound regressors for each participant. ICA-AROMA ICA-AROMA (automatic removal of motion artifact) is a recently-introduced, widely-used method for de-noising using single-subject ICA (Pruim et al., 2015a,b); we eval- uated ICA-AROMA in confound regression model 12. In contrast to other ICA based methods (e.g., ICA-FIX: Salimi-Khorshidi et al. (2014)), it does not require dataset- specific training data. The input to ICA-AROMA is a vox- elwise time series that has been smoothed at 6mm FWHM using a Gaussian kernel. After decomposing this time se- ries using FSL's melodic (with model order estimated us- ing the Laplace approximation) (Beckmann et al., 2005), ICA-AROMA uses four features to determine whether each component corresponds to signal or noise. The first 2 features are spatial characteristics of the signal source: (1) 4 the fraction of the source that falls within a CSF compart- ment and (2) the fraction of the source that falls along the edge or periphery of the brain. The remaining fea- tures are derived from the time series of the source: (3) its maximal robust correlation with time series derived from realignment parameters and (4) its high-frequency spec- tral content. ICA-AROMA includes two de-noising steps. The first de-noising step occurs immediately after classi- fication. All component time series (signal and noise) are included as predictors in the linear model, and the resid- ual BOLD time series is obtained via partial regression of only the noise time series. A second confound regression step occurs after temporal filtering, wherein mean signals from WM and CSF were regressed from the data. Temporal censoring: scrubbing de-spiking, spike regression, and In addition to regression of nuisance time series, a num- ber of 'temporal censoring' approaches were used to iden- tify motion-contaminated volumes in the BOLD time se- ries and reduce their impact on further analysis. These ap- proaches included despiking, spike regression, and scrub- bing. Despiking is a procedure that identifies outliers in the intensity of each voxel's detrended BOLD time series and then interpolates over these outliers. Despiking was implemented in AFNI using the 3dDespike utility (Cox, 1996) as part of confound regression model 6. Unlike despiking, which identifies outliers on a voxel- wise basis, spike regression and scrubbing censor complete volumes based on metrics of subject movement defined a priori. For spike regression, as in Satterthwaite et al. (2013), volumes were flagged for spike regression if their volume-to-volume RMS displacement exceeded 0.25mm. Next, as part of confound regression model 7, k 'spike' regressors were included as predictor variables in the de- noising model, where k equalled the number of volumes flagged (Satterthwaite et al., 2013). For each flagged time point, a unit impulse function that had a value of 1 at that time point and 0 elsewhere was included as a spike regressor. For scrubbing, the framewise displacement (FD) (Power et al., 2012) was computed at each time point as the sum of the absolute values of the derivatives of translational and rotational motion estimates. If framewise displace- ment (FD) at any point in time exceeded 0.2mm, then that time point was flagged for scrubbing. It should be noted that the conversion of FD to RMS displacement is approximately 2:1, and thus the published criterion for scrubbing has a lower threshold for flagging high-motion volumes than does spike regression. Scrubbing of BOLD data was performed iteratively (Power et al., 2014a) as part of confound regression model 8. At any stage where a linear model was applied to the data (for instance, dur- ing detrending procedures), high-motion epochs were tem- porally masked out of the model so as not to influence fit. During temporal filtering, a frequency transform was used to generate surrogate data with the same phase and spectral characteristics as the unflagged data. This sur- rogate data was used to interpolate over flagged epochs prior to application of the filter. During confound regres- sion, flagged timepoints were excised from the time series so as not to contribute to the model fit. For scrubbing (but not spike regression) if fewer than five contiguous volumes had unscrubbed data, these volumes were scrubbed and interpolated as well. Overview of outcome measures We evaluated each de-noising pipeline according to three benchmarks. Residual QC-FC correlations and distance- dependence provided a metric of each pipeline's efficacy, while loss of temporal DOF provided an estimate of each pipeline's efficiency. QC-FC correlations In order to estimate the residual relationship between subject movement and connectivity after de-noising, we computed QC-FC correlations (quality control / func- tional connectivity) (Power et al., 2015; Satterthwaite et al., 2012, 2013; Power et al., 2012). While other met- rics have been used in prior reports, including FD-DVARS correlations, we favor QC-FC as the most useful metric of interest as it directly quantifies the relationship between motion and the primary outcome of interest (rather than two quality metrics, as in FD-DVARS). For an extended discussion of the rationale for this measure, see Power et al. (2015). We evaluated QC-FC relationships within two commonly-used whole-brain networks, the first consisting of spherical nodes distributed across the brain (Power et al., 2011) and the second comprising an areal par- cellation of the cerebral cortex (Gordon et al., 2016). For each network, the mean time series for each node was calculated from the denoised residual data, and the pairwise Pearson correlation coefficient between node time series was used as the network edge weight (Biswal et al., 1995). For each edge, we then computed the correlation between the weight of that edge and the mean relative RMS motion. To eliminate the potential influence of demographic factors, QC-FC relationships were calculated as partial correlations that accounted for participant age and sex. We thus obtained, for each de-noising pipeline, a distribution of QC-FC correlations. This distribution was used to obtain two measures of the pipeline's ability to mitigate motion artifact, including: 1) the number of edges significantly related to motion, which was computed after using the false discovery rate (FDR) to account for multiple comparisons; and 2) the median absolute value of all QC-FC correlations. All graphs were generated using ggplot2 in R version 3.2.3; brain renderings were prepared in BrainNet Viewer (Xia et al., 2013). 5 Distance-dependent effects of motion Early work on motion artifact demonstrated that in- scanner motion can bias connectivity estimates between two nodes in a manner that is related to the distance between those nodes (Satterthwaite et al., 2012; Power et al., 2012). Under certain processing conditions, subject movement enhances short-distance connections while re- ducing long-distance connections. To determine the resid- ual distance-dependence of subject movement, we first used the center of mass of each node to obtain a distance matrix D where entry Dij indicates the Euclidean dis- tance between the centers of mass of nodes i and j. We then obtained the correlation between the distance sepa- rating each pair of nodes and the QC-FC correlation of the edge connecting those nodes; this correlation served as an estimate of the distance-dependence of motion artifact. Additional degrees of freedom lost in confound regression Including additional regressors in a confound model re- duces the impact of motion on future analyses, but it is not without cost. Confound regressors and censoring both reduce the temporal degrees of freedom (DOF) in data. This loss in temporal DOF may diminish the ability of functional data acquisitions to sample a subject's connec- tome or may introduce bias if the loss is variable across subjects. Thus, de-noising strategies ideally limit the loss of temporal DOF, for instance by including fewer, more efficacious regressors. In the present study, we also as- sessed the number of temporal DOF lost in each confound regression approach. As in previous work (Pruim et al., 2015a), we assumed that each time series regressed out and each volume ex- cised from the data constituted a single temporal DOF. Consequently, the loss of temporal DOF was estimated as the sum of the number of regressors in each confound model and the number of volumes flagged for excision un- der that model. It should be emphasized that the val- ues thus obtained are imperfect estimates. First, because functional MR time series typically exhibit temporal au- tocorrelation, the actual loss in DOF will be less than the estimated loss in DOF. Accordingly, censoring adjacent volumes does not remove the same number of DOF as does censoring volumes separated in time. Furthermore, a temporal bandpass filter was uniformly applied to all data prior to confound regression; this filtering procedure would itself have removed additional temporal DOF and elevated the autocorrelation of the data. Because this fil- ter was uniform across all de-noising strategies, it was not considered when estimating the loss of additional temporal DOF in each model. Results Heterogeneity in confound regression performance Confound regression strategies typically remove some, but not all, of the artifactual variance that head motion 6 introduces into the BOLD signal. The motion-related arti- fact that survives de-noising can be quantified to provide a metric of pipeline performance. Here, our primary bench- mark of confound regression efficacy was the residual rela- tionship between brain connectivity and subject motion, or the QC-FC correlation. We measured QC-FC correlations using two metrics: the percentage of network connections where a significant relationship with motion was present (Figure 2), and the absolute median correlation (AMC) between connection strength and head movement across all connections (Figure 3). No preprocessing strategy was completely effective in abolishing the relationship between head movement and connectivity. However, different approaches exhibited widely varying degrees of efficacy. The top four con- found regression strategies included 36 parameters, com- prising an expansion of GSR, tissue-specific regressors (WM, CSF), and realignment parameters. Beyond this base 36-parameter model, all censoring techniques pro- vided provided some additional benefit, reducing the num- ber of edges that were significantly related to motion to less than 1%. Convergent results were present across both QC-FC measures (% edges, AMC) and networks (Power, Gordon) that were evaluated. In contrast, many pipelines performed relatively poorly, leaving a majority of network edges with a residual re- lationship with motion. Specifically, 82% of edges were impacted by motion when the least effective method was used (6 realignment parameters). The commonly used 24- parameter expansion of realignment parameters originally suggested by Friston et al. (1996) did not provide much of an improvement (79% edges). Similarly, the local WM regressor model (69% edges) and tCompCor (48% edges) also resulted in substantial residual QC-FC correlations. In fact, these methods performed worse than a basic 2- parameter model composed of mean WM and CSF signals (33% edges). Finally, several methods demonstrated in- termediate performance, with 1-20% of edges impacted by motion. This middle group included methods as disparate as aCompCor (5% edges), ICA-AROMA (12% edges), and the classic 9-parameter confound regression model which included GSR (4% edges). Variability in distance-dependent motion artifact after confound regression Our second benchmark quantified the distance- dependent motion artifact that was present in data pro- cessed by each pipeline (Figure 4). We observed that distance-dependence was present even under conditions where artifact magnitude was attenuated. For example, though the 36-parameter model was among the most ef- fective in attenuating QC-FC relationships, its application revealed strongly distance-dependent artifact. Examina- tion of graphs that plot QC-FC by Euclidean distance (see Figure 4C) revealed that this is due to effective mitiga- tion of motion artifact for long-range but not short-range connections. Distance-dependence was most prominent in models that included GSR, but did not include censoring (e.g., 9-parameter and 36-parameter models). However, despite the lack of global signal in the aCompCor or tCompcor models, data returned from both of these component- based approaches revealed substantial distance-dependent artifact. Notably, inclusion of censoring consistently re- duced distance-dependence, although scrubbing was some- what more effective than spike regression or voxelwise de- spiking. The top performing method according to this bench- mark was ICA-AROMA, which completely abolished any distance-dependence of residual motion artifact. In other words, the motion artifact that was still present in the data after ICA-AROMA impacted all connections in a manner that was not dependent on the spatial separation between nodes. There was similar lack of distance-dependence in the wmLocal model, although as noted above this model did not provide effective de-noising according to QC-FC benchmarks. Despite the presence of the global signal, the 36-parameter model that included scrubbing also per- formed quite well. Effective preprocessing strategies use many additional de- grees of freedom Perhaps unsurprisingly, the preprocessing strategies that consistently reduced both QC-FC correlations and distance-dependence were also among the costliest in terms of loss of temporal degrees of freedom (Figure 5). By definition, the 36-parameter models included a high fixed number of regressors. Furthermore, models that additionally included censoring resulted in a substantial additional loss of data that varied across subjects. ICA- AROMA also had a variable loss of DOF, but of a lower magnitude than censoring or high-parameter confound re- gression. Discussion In response to rapid evolution of confound regression strategies available for the mitigation of motion artifact, in this report we evaluated 12 commonly-used pipelines. Results indicate that there is substantial heterogeneity in the performance of these confound regression techniques across all measures evaluated. The context, implications, and limitations of these findings are discussed below. Confound regression techniques have substantial perfor- mance variability We evaluated confound regression strategies according to three intuitive benchmarks that were selected to capture different domains of effectiveness. These included QC-FC associations, distance-dependence of motion artifact, and additional degrees of freedom lost in confound regression. Across each benchmark, there was a striking heterogeneity in pipeline performance. Notably, five of the top six confound regression ap- proaches included GSR. This effect was consistent in both networks we evaluated. The effectiveness of GSR is most likely due to the nature of motion artifact itself: in-scanner head motion tends to induce widespread reductions in sig- nal intensity across the entire brain parenchyma (see Sat- terthwaite et al. (2013), Figure 4). As discussed in de- tail elsewhere (Power et al., Under Revision) this effect is highly reproducible across datasets, and is effectively captured by time series regression of the global signal. Beyond GSR, a second strategy that clearly minimizes QC-FC relationships is temporal censoring. We evaluated three censoring variants, including scrubbing, spike regres- sion, and de-spiking. Compared to spike regression and de-spiking, scrubbing appears to be more effective in re- moving distance-dependent artifact in this dataset. This is most likely due to the explicit tension between data qual- ity and data quantity: because of the lower threshold for scrubbing than spike regression (due to differences in FD vs. RMS measures of motion; see Figure 9C in Yan et al. (2013a)), more low-quality data was excised during scrub- bing. Furthermore, scrubbing includes a criterion to not leave isolated epochs (< 5 volumes) of un-scrubbed data. Consequently, this leads to clear differences in the addi- tional degrees of freedom lost by each method. In contrast to spike regression and scrubbing, which eliminate high motion volumes completely, time series de-spiking identi- fies and interpolates large changes in signal intensity on a voxelwise basis (Cox, 1996). This allows for spatial adap- tivity (see Patel et al. (2014)) but also renders quantifi- cation of data loss and comparisons with volume-based censoring techniques more difficult. Critically, while both GSR and censoring appeared ef- fective in minimizing QC-FC relationships, they exhibited opposite effects on distance-dependence. While censor- ing techniques appear to consistently reduce the presence of distance-dependence, GSR is associated with increased distance-dependence. However, it should be emphasized that the distance-dependence associated with GSR is not the result of worsening associations with motion in certain connections. Rather, the distance-dependence seen with GSR stems from differential de-noising efficacy, whereby motion artifact is more effectively minimized for long- distance connections than for short-range connections. Certain models such as the local WM regression approach (Jo et al., 2013) thus have minimal distance-dependence, but this is a consequence of lack of efficacy across all dis- tances. In contrast, ICA-AROMA (Pruim et al., 2015b,a) reduced motion to a moderate degree over all connec- tion distances, resulting in almost no distance-dependence. However, while clearly an improvement over some other methods, data processed using ICA-AROMA was nois- ier than other methods which included GSR or censor- ing, and resulting networks contained a substantial num- ber of edges impacted by motion. Somewhat to our sur- prise, benchmark results for aCompCor (Behzadi et al., 2007; Muschelli et al., 2014) were most similar to models 7 that included GSR. Alone among models where GSR was not included, aCompCor was both relatively effective in the mitigation of residual motion (5% of edges impacted) and also exhibited substantial distance-dependence (e.g., r = −0.26). This suggests that while aCompCor does not explicitly include GSR, the practical results of its applica- tion are in fact quite similar. Trade-offs of confound regression approaches: implications for investigators The current results emphasize two clear trade-offs in the choice of confound regression strategy. First, pipelines that include global signal regression tend to be more effec- tive at minimizing QC-FC relationships, but at the cost of some increase in distance-dependence. As noted above, for minimizing QC-FC relationships, nearly all of the top strategies (except aCompCor) included GSR. Conversely, the two techniques that had the most substantial distance- dependence (the 9-regressor and 36-regressor methods) both included GSR. Second, censoring techniques provide a clear benefit in reducing QC-FC relationships and addi- tionally tend to attenuate distance-dependence. However, by definition, removing contaminated volumes results in less data and loss of degrees of freedom. These two trade-offs suggest that a single confound re- gression strategy is unlikely to be optimal for every study. For example, in studies of network organization, the pres- ence of both anti-correlations and distance-dependent ar- tifact resulting from GSR may result in an undesirable impact on nodal degree distribution (Yan et al., 2013b). In contrast, for studies of group or individual differ- ences, minimizing QC-FC relationships is likely to be of paramount importance so as to limit the possibility that inference is driven by artifactual signals. This concern is particularly relevant for studies of development or clinical sub-groups where motion is systematically related to the subject-level variable of interest (e.g., age, disease status). Co-varying for motion at the group level is unlikely to be a panacea for such studies when inadequate subject-level time series de-noising is employed, as prior work (Power et al., 2014a) has suggested that motion effects at the group level may potentially both be nonlinear and vary across sub-samples in a manner that is difficult to predict. However, aggressive volume censoring may be problematic in datasets with relatively brief acquisitions. In datasets where long time series are acquired, such as multi-band ac- quisitions (Feinberg et al., 2010) and intensive acquisitions of single subjects (Laumann et al., 2015), loss of tempo- ral degrees of freedom is less likely to be a major concern. The 36-parameter models without volume censoring offer uniformity, as does randomly or systematically censoring additional volumes until all subjects retain approximately the same degrees of freedom. Limitations Several limitations of the current approach should be noted. One of the principal challenges in evaluating the performance of de-noising approaches is the lack of a noise- free ground truth. Our primary benchmark of confound regression performance assumes that mitigation of the re- lationship between QC (participant motion) and FC (i.e., the imaging measurement) is desirable. To the degree that in-scanner motion itself represents a biologically informa- tive phenotype, this approach will mistake signal for noise. Indeed, prior data suggests that this may sometimes be the case. For example, Zeng et al. (2014) found specific changes in connectivity for participants who had generally high levels of motion, even on scans where motion was low. However, without multiple scans which allow such careful dissociation, most studies are incapable of disambiguating the large confounding effects of motion on connectivity. Second, in place of QC-FC relationships, one could focus on alternative benchmarks such as test-retest reliability (Zuo et al., 2014). Reliability is certainly of interest, but to the degree that motion tends to be highly correlated within individuals across scanning sessions, there is a sub- stantial potential for the presence of consistent motion ar- tifact across sessions to artificially inflate reliability, and diminish the biological relevance of observed results. A third and related concern is that certain de-noising meth- ods could conceivably both minimize QC-FC relationships and even enhance reliability by aggressively removing both signal and noise, but in the process diminish sensitivity to meaningful individual differences. Indeed, one prior study demonstrated the association between canonical resting state networks and randomly generated confound parame- ters (Bright and Murphy, 2015). This concern is somewhat mitigated by other studies in independent datasets, which suggest that higher-order confound regressors improve the confound regression model fit (Yan et al., 2013a; Satterth- waite et al., 2013), while random regressors do not (see Fig- ure 8 in Satterthwaite et al. (2013)). Nonetheless, the ten- sion between the goals of noise reduction and sensitivity to individual differences remains an outstanding issue for the field. Fourth, while our evaluation included many of the most commonly used techniques, other approaches which require substantial training or parameter selection (i.e., ICA-FIX (Salimi-Khorshidi et al., 2014; Griffanti et al., 2014), wavelet de-spiking (Patel et al., 2014)) may be valu- able and merit further consideration. Fifth and finally, it should be noted that while improvements in image acqui- sition (including multi-echo techniques) may not salvage existing motion-contaminated data, it is likely that they will change the methodological landscape of connectivity research moving forward (Kundu et al., 2012, 2013; Bright and Murphy, 2013). Conclusions Taken together, the present results underline the per- formance heterogeneity of recently-introduced, commonly- used confound regression methods. In selecting among these methods, investigators should be aware of the rel- ative strengths and weaknesses of each approach, and un- derstand how processing strategy may impact inference. 8 Clearly, the relative merit of each approach will vary by research question and study design. Perhaps most im- portantly, as has been emphasized in nearly every other study of motion artifact, the choice of confound regres- sion strategy is often dwarfed in importance by the need to transparently report and evaluate the impact of mo- tion in each dataset. At a minimum, this includes report- ing the relationship between motion artifact and not only subject phenotypes (e.g., group, age, symptom or cogni- tive score) but also the functional connectivity phenotypes being considered. In the context of such data, the distinc- tion between observed results and the impact of motion artifact can be understood. Such transparency bolsters confidence in reported findings, but also will likely tend to emphasize the remaining challenges for de-noising go- ing forward. Especially when considered in the context of the the rapid evolution of available techniques since 2012, there is no doubt that innovations in post-processing con- found regression strategies will continue. Acknowledgements Thanks to the acquisition and recruitment team, in- cluding Karthik Prabhakaran and Jeff Valdez. Thanks to Chad Jackson for data management and systems support. Thanks to Monica Calkins for phenotyping expertise. Sup- ported by grants from the National Institute of Men- tal Health: R01MH107703 (TDS), R01MH107235 (RCG), and R01NS089630 (CD). The PNC was funded through NIMH RC2 grants MH089983, and MH089924 (REG). Additional support was provided by R21MH106799 (DSB & TDS), R01MH101111 (DHW), K01MH102609 (DRR), P50MH096891 (REG), R01NS085211 (RTS), and the Dowshen Program for Neuroscience. DSB acknowledges support from the John D. and Catherine T. MacArthur Foundation, the Alfred P. Sloan Foundation, the Army Re- search Laboratory and the Army Research Office through contract numbers W911NF-10-2-0022 and W911NF-14- 1-0679, the National Institute of Mental Health (R01- DC-009209-11), the National Institute of Child Health and Human Development (R01HD086888-01), the Of- fice of Naval Research, and the National Science Foun- dation (BCS-1441502 and PHY-1554488). Support for developing statistical analyses (RTS & TDS) was pro- vided by a seed grant by the Center for Biomedi- cal Computing and Image Analysis (CBICA) at Penn. Data deposition: The data reported in this paper have been deposited in database of Genotypes and Pheno- types (dbGaP), www.ncbi.nlm.nih.gov/gap (accession no. phs000607.v1.p1). References References Avants, B.B., Tustison, N.J., Song, G., Cook, P.A., Klein, A., Gee, J.C., 2011a. A reproducible evaluation of ANTs similarity metric 9 performance in brain image registration. NeuroImage 54, 2033– 2044. doi:10.1016/j.neuroimage.2010.09.025. Avants, B.B., Tustison, N.J., Wu, J., Cook, P.A., Gee, J.C., 2011b. An open source multivariate framework for N-tissue segmenta- tion with evaluation on public data. Neuroinformatics 9, 381–400. doi:10.1007/s12021-011-9109-y. Baker, J.T., Holmes, A.J., Masters, G.a., Yeo, B.T.T., Krienen, F., Buckner, R.L., Ongur, D., 2014. Disruption of cortical association networks in schizophrenia and psychotic bipolar disorder. JAMA psychiatry 71, 109–18. doi:10.1001/jamapsychiatry.2013.3469, arXiv:15334406. Beckmann, C.F., Deluca, M., Devlin, J.T., Smith, S.M., 2005. Inves- tigations into resting-state connectivity using independent compo- nent analysis. Philos Trans R Soc Lond B Biol Sci 360, 1001–13. doi:10.1098/rstb.2005.1634. Behzadi, Y., Restom, K., Liau, J., Liu, T.T., 2007. A compo- nent based noise correction method (CompCor) for BOLD and perfusion based fMRI. NeuroImage 37, 90–101. doi:10.1016/j. neuroimage.2007.04.042, arXiv:NIHMS150003. Biswal, B., Yetkin, F.Z., Haughton, V.M., Hyde, J.S., 1995. Func- tional connectivity in the motor cortex of resting human brain using echo-planar MRI. Magnetic Resonance in Medicine 34, 537– 541. doi:10.1002/mrm.1910340409. Bright, M.G., Murphy, K., 2013. Removing motion and physiolog- ical artifacts from intrinsic BOLD fluctuations using short echo data. NeuroImage 64, 526–537. doi:10.1016/j.neuroimage.2012. 09.043. Bright, M.G., Murphy, K., 2015. Is fMRI "noise" really noise? Resting state nuisance regressors remove variance with network structure. NeuroImage 114, 158–169. doi:10.1016/j.neuroimage. 2015.03.070. Buckner, R.L., Andrews-Hanna, J.R., Schacter, D.L., 2008. The brain's default network: Anatomy, function, and relevance to dis- ease. Annals of the New York Academy of Sciences 1124, 1–38. doi:10.1196/annals.1440.011. Chai, X.J., Castan´an, A.N., Ongur, D., Whitfield-Gabrieli, S., 2012. Anticorrelations in resting state networks without global signal re- gression. NeuroImage 59, 1420–1428. doi:10.1016/j.neuroimage. 2011.08.048. Cox, R., 1996. AFNI: Software for analysis and visualization of functional magnetic resonance neuroimages. Comput Biomed Res 29, 162–173. Craddock, R.C., Jbabdi, S., Yan, C.G., Vogelstein, J.T., Castel- lanos, F.X., Di Martino, A., Kelly, C., Heberlein, K., Colcombe, S., Milham, M.P., 2013. Imaging human connectomes at the macroscale. Nature Methods 10, 524–539. doi:10.1038/nmETh. 2482, arXiv:NIHMS150003. Damoiseaux, J.S., Rombouts, S.A.R.B., Barkhof, F., Scheltens, P., Stam, C.J., Smith, S.M., Beckmann, C.F., 2006. Consistent resting-state networks across healthy subjects. Proceedings of the National Academy of Sciences of the United States of America 103, 13848–53. doi:10.1073/pnas.0601417103. DiMartino, A., Fair, D.A., Kelly, C., Satterthwaite, T.D., Castel- lanos, F.X., Thomason, M.E., Craddock, R.C., Luna, B., Lev- enthal, B.L., Zuo, X.N., Milham, M.P., 2014. Unraveling the miswired connectome: A developmental perspective. Neuron 83, 1335–1353. doi:10.1016/j.neuron.2014.08.050. Dosenbach, N.U.F., Nardos, B., Cohen, A.L., Fair, D.A., Power, D., Church, J.a., Nelson, S.M., Wig, G.S., Vogel, A.C., Lessov- schlaggar, C.N., Barnes, K.A., Dubis, J.W., Feczko, E., Coalson, R.S., Jr, J.R.P., Barch, D.M., Petersen, S.E., Schlaggar, B.L., 2011. Prediction of individual brain maturity using fMRI. Science 329, 1358–1361. doi:10.1126/science.1194144.Prediction. Fair, D.A., Cohen, A.L., Dosenbach, N.U.F., Church, J.A., Miezin, F.M., Barch, D.M., Raichle, M.E., Petersen, S.E., Schlaggar, B.L., 2008. The maturing architecture of the brain's default network. Proceedings of the National Academy of Sciences of the United States of America 105, 4028–4032. doi:10.1073/pnas.0800376105. Fair, D.A., Nigg, J.T., Iyer, S., Bathula, D., Mills, K.L., Dosenbach, N.U.F., Schlaggar, B.L., Mennes, M., Gutman, D., Bangaru, S., Buitelaar, J.K., Dickstein, D.P., Di Martino, A., Kennedy, D.N., Kelly, C., Luna, B., Schweitzer, J.B., Velanova, K., Wang, Y.F., Mostofsky, S., Castellanos, F.X., Milham, M.P., 2012. Distinct neural signatures detected for ADHD subtypes after controlling for micro-movements in resting state functional connectivity MRI data. Frontiers in systems neuroscience 6, 80. doi:10.3389/fnsys. 2012.00080. Fair, D.A., Posner, J., Nagel, B.J., Bathula, D., Dias, T.G.C., Mills, K.L., Blythe, M.S., Giwa, A., Schmitt, C.F., Nigg, J.T., 2010. Atypical default network connectivity in youth with attention- deficit/hyperactivity disorder. Biological psychiatry 68, 1084–91. doi:10.1016/j.biopsych.2010.07.003. Feinberg, D.A., Moeller, S., Smith, S.M., Auerbach, E.J., Ramanna, S., Glasser, M.F., Miller, K.L., Ugurbil, K., Yacoub, E.S., 2010. Multiplexed echo planar imaging for sub-second whole brain fMRI and fast diffusion imaging. PLoS ONE 5, e15710. doi:10.1371/ journal.pone.0015710. Fox, M.D., Raichle, M.E., 2007. Spontaneous fluctuations in brain activity observed with functional magnetic resonance imaging. Nat Rev Neurosci 8, 700–711. doi:10.1038/nrn2201. Fox, M.D., Snyder, A.Z., Vincent, J.L., Corbetta, M., Van Es- sen, D.C., Raichle, M.E., 2005. The human brain is intrinsi- cally organized into dynamic, anticorrelated functional networks. Proceedings of the National Academy of Sciences of the United States of America 102, 9673–8. doi:10.1073/pnas.0504136102, arXiv:NIHMS150003. Fox, M.D., Zhang, D., Snyder, A.Z., Raichle, M.E., 2009. The global signal and observed anticorrelated resting state brain net- works. Journal of neurophysiology 101, 3270–3283. doi:10.1152/ jn.90777.2008. Friston, K., Williams, S., Howard, R., Frackowiak, R.S.J., Turner, R., 1996. Movement-related effects in {fMRI} time-series. Mag- netic Resonance in Medicine 35, 346–355. Gordon, E.M., Laumann, T.O., Adeyemo, B., Huckins, J.F., Kel- ley, W.M., Petersen, S.E., 2016. Generation and evaluation of a cortical area parcellation from resting-state correlations. Cerebral Cortex 26, 288–303. doi:10.1093/cercor/bhu239. Gotts, S.J., Saad, Z.S., Jo, H.J., Wallace, G.L., Cox, R.W., Martin, A., 2013. The perils of global signal regression for group compar- isons: a case study of Autism Spectrum Disorders. Frontiers in Human Neuroscience 7, 356. doi:10.3389/fnhum.2013.00356. Greve, D.N., Fischl, B., 2009. Accurate and robust brain image alignment using boundary-based registration. NeuroImage 48, 63– 72. doi:10.1016/j.neuroimage.2009.06.060. Griffanti, L., Salimi-Khorshidi, G., Beckmann, C.F., Auerbach, E.J., Douaud, G., Sexton, C.E., Zsoldos, E., Ebmeier, K.P., Filippini, N., Mackay, C.E., Moeller, S., Xu, J., Yacoub, E., Baselli, G., Ugurbil, K., Miller, K.L., Smith, S.M., 2014. ICA-based artefact removal and accelerated fMRI acquisition for improved resting state network imaging. NeuroImage 95, 232–247. doi:10.1016/j. neuroimage.2014.03.034. Gur, R.E., Kaltman, D., Melhem, E.R., Ruparel, K., Prabhakaran, K., Riley, M., Yodh, E., Hakonarson, H., Satterthwaite, T., Gur, R.C., 2013. Incidental findings in youths volunteering for brain MRI research. American Journal of Neuroradiology 34, 2021–2025. doi:10.3174/ajnr.A3525. Hahamy, A., Calhoun, V., Pearlson, G., Harel, M., Stern, N., Attar, F., Malach, R., Salomon, R., 2014. Save the global: global signal connectivity as a tool for studying clinical populations with func- tional magnetic resonance imaging. Brain connectivity 4, 395–403. doi:10.1089/brain.2014.0244. Hallquist, M.N., Hwang, K., Luna, B., 2013. The nuisance of nuisance regression: Spectral misspecification in a common ap- proach to resting-state fMRI preprocessing reintroduces noise and obscures functional connectivity. NeuroImage 82, 208–225. doi:10.1016/j.neuroimage.2013.05.116, arXiv:NIHMS150003. Jenkinson, M., Bannister, P., Brady, M., Smith, S., 2002. the robust and accurate linear Neu- doi:10.1016/S1053-8119(02)91132-8, Improved optimization for registration and motion correction of brain images. roImage 17, 825–841. arXiv:arXiv:1011.1669v3. Cox, R.W., Saad, Z.S., 2013. Effective preprocessing procedures virtually eliminate distance-dependent motion artifacts in resting state FMRI. Journal of Applied Mathematics 2013. doi:10.1155/ 2013/935154. Klein, A., Andersson, J., Ardekani, B.A., Ashburner, J., Avants, B., Chiang, M.C., Christensen, G.E., Collins, D.L., Gee, J., Hel- lier, P., Song, J.H., Jenkinson, M., Lepage, C., Rueckert, D., Thompson, P., Vercauteren, T., Woods, R.P., Mann, J.J., Parsey, R.V., 2009. Evaluation of 14 nonlinear deformation algorithms ap- plied to human brain MRI registration. NeuroImage 46, 786–802. doi:10.1016/j.neuroimage.2008.12.037. Kundu, P., Brenowitz, N.D., Voon, V., Worbe, Y., V´ertes, P.E., Inati, S.J., Saad, Z.S., Bandettini, P.A., Bullmore, E.T., 2013. Integrated strategy for improving functional connectivity mapping using multiecho fMRI. Proceedings of the National Academy of Sciences of the Unites States of America 110, 16187–16192. doi:10. 1073/pnas.1301725110, arXiv:arXiv:1408.1149. Kundu, P., Inati, S.J., Evans, J.W., Luh, W.M., Bandettini, P.A., 2012. Differentiating BOLD and non-BOLD signals in fMRI time series using multi-echo EPI. NeuroImage 60, 1759–1770. doi:10. 1016/j.neuroimage.2011.12.028. Laumann, T.O., Gordon, E.M., Adeyemo, B., Snyder, A.Z., Joo, S.J., Chen, M.Y., Gilmore, A.W., McDermott, K.B., Nelson, S.M., Dosenbach, N.U.F., Schlaggar, B.L., Mumford, J.A., Pol- drack, R.A., Petersen, S.E., 2015. Functional system and areal organization of a highly sampled individual human brain. Neuron 87, 658–671. doi:10.1016/j.neuron.2015.06.037. Macey, P.M., Macey, K.E., Kumar, R., Harper, R.M., 2004. A method for removal of global effects from fMRI time series. Neu- roImage 22, 360–366. doi:10.1016/j.neuroimage.2003.12.042. Merikangas, K.R., He, J.P., Brody, D., Fisher, P., Bourdon, K., Ko- retz, D.S., 2010. Prevalence and treatment of mental disorders among US children in the 20012004 NHANES. Pediatrics 125, 75–81. doi:10.1542/peds.2008-2598.Prevalence. Murphy, K., Birn, R.M., Handwerker, D.A., Jones, T.B., Bandet- tini, P.A., 2009. The impact of global signal regression on resting state correlations: Are anti-correlated networks introduced? Neu- roImage 44, 893–905. doi:10.1016/j.neuroimage.2008.09.036, arXiv:NIHMS150003. Muschelli, J., Nebel, M.B., Caffo, B.S., Barber, A.D., Pekar, J.J., Mostofsky, S.H., 2014. Reduction of motion-related artifacts in resting state fMRI using aCompCor. NeuroImage 96, 22–35. doi:10.1016/j.neuroimage.2014.03.028. Patel, A.X., Kundu, P., Rubinov, M., Jones, P.S., V´ertes, P.E., Ersche, K.D., Suckling, J., Bullmore, E.T., 2014. A wavelet method for modeling and despiking motion artifacts from resting- state fMRI time series. NeuroImage 95, 287–304. doi:10.1016/j. neuroimage.2014.03.012. Power, J.D., Barnes, K.A., Snyder, A.Z., Schlaggar, B.L., Petersen, S.E., 2012. Spurious but systematic correlations in functional connectivity MRI networks arise from subject motion. Neu- roImage 59, 2142–2154. doi:10.1016/j.neuroimage.2011.10.018, arXiv:NIHMS150003. Power, J.D., Cohen, A.L., Nelson, S.M., Wig, G.S., Barnes, K.A., Church, J.A., Vogel, A.C., Laumann, T.O., Miezin, F.M., Schlag- gar, B.L., Petersen, S.E., 2011. Functional network organization of the human brain. Neuron 72, 665–678. doi:10.1016/j.neuron. 2011.09.006. Power, J.D., Mitra, A., Laumann, T.O., Snyder, A.Z., Schlag- gar, B.L., Petersen, S.E., 2014a. Methods to detect, charac- terize, and remove motion artifact in resting state fMRI. Neu- roImage 84, 320–341. doi:10.1016/j.neuroimage.2013.08.048, arXiv:NIHMS150003. Power, J.D., Schlaggar, B.L., Petersen, S.E., 2014b. Studying brain organization via spontaneous fMRI signal. Neuron 84, 681–696. doi:10.1016/j.neuron.2014.09.007, arXiv:NIHMS150003. Power, J.D., Schlaggar, B.L., Petersen, S.E., 2015. Recent progress and outstanding issues in motion correction in resting state fMRI. NeuroImage 105, 536–551. doi:10.1016/j.neuroimage.2014.10. 044, arXiv:NIHMS150003. Jo, H.J., Gotts, S.J., Reynolds, R.C., Bandettini, P.A., Martin, A., Pruim, R.H.R., Mennes, M., Buitelaar, J.K., Beckmann, C.F., 2015a. 10 visualization tool for human brain connectomics. PLoS ONE 8. doi:10.1371/journal.pone.0068910. Yan, C.G., Cheung, B., Kelly, C., Colcombe, S., Craddock, R.C., Di Martino, A., Li, Q., Zuo, X.N., Castellanos, F.X., Milham, M.P., 2013a. A comprehensive assessment of regional variation in the impact of head micromovements on functional connectomics. Neu- roImage 76, 183–201. doi:10.1016/j.neuroimage.2013.03.004, arXiv:NIHMS150003. Yan, C.G., Craddock, R.C., He, Y., Milham, M.P., 2013b. Ad- dressing head motion dependencies for small-world topologies in functional connectomics. Frontiers in human neuroscience 7, 910. doi:10.3389/fnhum.2013.00910. Yan, C.G., Craddock, R.C., Zuo, X.N., Zang, Y.F., Milham, M.P., 2013c. Standardizing the intrinsic brain: Towards robust measure- ment of inter-individual variation in 1000 functional connectomes. NeuroImage 80, 246–262. doi:10.1016/j.neuroimage.2013.04. 081, arXiv:NIHMS150003. Yang, G.J., Murray, J.D., Repovs, G., Cole, M.W., Savic, A., Glasser, M.F., Pittenger, C., Krystal, J.H., Wang, X.J., Pearl- son, G.D., Glahn, D.C., Anticevic, A., 2014. Altered global brain signal in schizophrenia. Proceedings of the National Academy of Sciences of the United States of America 111, 7438–43. doi:10. 1073/pnas.1405289111. Yeo, B.T., Krienen, F.M., Sepulcre, J., Sabuncu, M.R., Lashkari, D., Hollinshead, M., Roffman, J.L., Smoller, J.W., Zollei, L., Poli- meni, J.R., Fischl, B., Liu, H., Buckner, R.L., 2011. The organi- zation of the human cerebral cortex estimated by intrinsic func- tional connectivity. Journal of neurophysiology 106, 1125–1165. doi:10.1152/jn.00338.2011. Zeng, L.L., Wang, D., Fox, M.D., Sabuncu, M., Hu, D., Ge, M., Buckner, R.L., Liu, H., 2014. Neurobiological basis of head motion in brain imaging. Proceedings of the National Academy of Sciences of the United States of America 111, 6058–62. doi:10.1073/pnas. 1317424111. Zuo, X.N., Anderson, J.S., Bellec, P., Birn, R.M., Biswal, B.B., Blautzik, J., Breitner, J.C.S., Buckner, R.L., Calhoun, V.D., Castellanos, F.X., Chen, A., Chen, B., Chen, J., Chen, X., Col- combe, S.J., Courtney, W., Craddock, R.C., Di Martino, A., Dong, H.M., Fu, X., Gong, Q., Gorgolewski, K.J., Han, Y., He, Y., He, Y., Ho, E., Holmes, A., Hou, X.H., Huckins, J., Jiang, T., Jiang, Y., Kelley, W., Kelly, C., King, M., LaConte, S.M., Lain- hart, J.E., Lei, X., Li, H.J., Li, K., Li, K., Lin, Q., Liu, D., Liu, J., Liu, X., Liu, Y., Lu, G., Lu, J., Luna, B., Luo, J., Lurie, D., Mao, Y., Margulies, D.S., Mayer, A.R., Meindl, T., Meyerand, M.E., Nan, W., Nielsen, J.A., O'Connor, D., Paulsen, D., Prabhakaran, V., Qi, Z., Qiu, J., Shao, C., Shehzad, Z., Tang, W., Villringer, A., Wang, H., Wang, K., Wei, D., Wei, G.X., Weng, X.C., Wu, X., Xu, T., Yang, N., Yang, Z., Zang, Y.F., Zhang, L., Zhang, Q., Zhang, Z., Zhang, Z., Zhao, K., Zhen, Z., Zhou, Y., Zhu, X.T., Milham, M.P., 2014. An open science resource for establishing re- liability and reproducibility in functional connectomics. Scientific data 1, 140049. doi:10.1038/sdata.2014.49. Evaluation of ICA-AROMA and alternative strategies for motion artifact removal in resting state fMRI. NeuroImage 112, 278–287. doi:10.1016/j.neuroimage.2015.02.063. Pruim, R.H.R., Mennes, M., van Rooij, D., Llera, A., Buitelaar, J.K., Beckmann, C.F., 2015b. ICA-AROMA: A robust ICA-based strat- egy for removing motion artifacts from fMRI data. NeuroImage 112, 267–277. doi:10.1016/j.neuroimage.2015.02.064. Saad, Z.S., Gotts, S.J., Murphy, K., Chen, G., Jo, H.J., Martin, A., Cox, R.W., 2012. Trouble at rest: how correlation patterns and group differences become distorted after global signal regression. Brain connectivity 2, 25–32. doi:10.1089/brain.2012.0080. Salimi-Khorshidi, G., Douaud, G., Beckmann, C.F., Glasser, M.F., Griffanti, L., Smith, S.M., 2014. Automatic denoising of func- tional MRI data: Combining independent component analysis and hierarchical fusion of classifiers. NeuroImage 90, 449–468. doi:10.1016/j.neuroimage.2013.11.046, arXiv:NIHMS150003. Satterthwaite, T.D., Connolly, J.J., Ruparel, K., Calkins, M.E., Jackson, C., Elliott, M.A., Roalf, D.R., Hopsona, R., Prab- hakaran, K., Behr, M., Qiu, H., Mentch, F.D., Chiavacci, R., Sleiman, P.M.A., Gur, R.C., Hakonarson, H., Gur, R.E., 2016. The Philadelphia Neurodevelopmental Cohort: A publicly avail- able resource for the study of normal and abnormal brain devel- opment in youth. NeuroImage 124, 1115–1119. doi:10.1016/j. neuroimage.2015.03.056. Satterthwaite, T.D., Elliott, M.A., Gerraty, R.T., Ruparel, K., Loug- head, J., Calkins, M.E., Eickhoff, S.B., Hakonarson, H., Gur, R.C., Gur, R.E., Wolf, D.H., 2013. An improved framework for confound regression and filtering for control of motion arti- fact in the preprocessing of resting-state functional connectivity data. NeuroImage 64, 240–256. doi:10.1016/j.neuroimage.2012. 08.052, arXiv:NIHMS150003. Satterthwaite, T.D., Elliott, M.A., Ruparel, K., Loughead, J., Prab- hakaran, K., Calkins, M.E., Hopson, R., Jackson, C., Keefe, J., Riley, M., Mentch, F.D., Sleiman, P., Verma, R., Davatzikos, C., Hakonarson, H., Gur, R.C., Gur, R.E., 2014. Neuroimaging of the Philadelphia Neurodevelopmental Cohort. NeuroImage 86, 544–553. doi:10.1016/j.neuroimage.2013.07.064. Satterthwaite, T.D., Wolf, D.H., Loughead, J., Ruparel, K., Elliott, M.A., Hakonarson, H., Gur, R.C., Gur, R.E., 2012. Impact of in- scanner head motion on multiple measures of functional connec- tivity: Relevance for studies of neurodevelopment in youth. Neu- roImage 60, 623–632. doi:10.1016/j.neuroimage.2011.12.063, arXiv:NIHMS150003. Smith, S.M., Vidaurre, D., Beckmann, C.F., Glasser, M.F., Jenk- inson, M., Miller, K.L., Nichols, T.E., Robinson, E.C., Salimi- Khorshidi, G., Woolrich, M.W., Barch, D.M., U??urbil, K., Van Essen, D.C., 2013. Functional connectomics from resting-state fMRI. Trends in Cognitive Sciences 17, 666–682. doi:10.1016/j. tics.2013.09.016. Tustison, N.J., Avants, B.B., Cook, P.A., Zheng, Y., Egan, A., Yushkevich, P.A., Gee, J.C., 2010. N4ITK: Improved N3 bias cor- rection. IEEE Transactions on Medical Imaging 29, 1310–1320. doi:10.1109/TMI.2010.2046908. Tustison, N.J., Cook, P.A., Klein, A., Song, G., Das, S.R., Duda, J.T., Kandel, B.M., van Strien, N., Stone, J.R., Gee, J.C., Avants, B.B., 2014. Large-scale evaluation of ANTs and FreeSurfer cortical thickness measurements. NeuroImage 99, 166–179. doi:10.1016/ j.neuroimage.2014.05.044. Van Dijk, K.R.A., Hedden, T., Venkataraman, A., Evans, K.C., Intrinsic functional con- Lazar, S.W., Buckner, R.L., 2010. nectivity as a tool for human connectomics: theory, proper- ties, and optimization. Journal of neurophysiology 103, 297–321. doi:10.1152/jn.00783.2009. Van Dijk, K.R.A., Sabuncu, M.R., Buckner, R.L., 2012. The in- fluence of head motion on intrinsic functional connectivity MRI. NeuroImage 59, 431–438. doi:10.1016/j.neuroimage.2011.07. 044, arXiv:NIHMS150003. Wang, H., Cao, Y., Syeda-mahmood, T., 2014. Multi-atlas segmen- tation with learning-based label fusion. Machine Learning in Med- ical Imaging 35, 256–263. doi:10.1007/978-3-319-10581-9_32. Xia, M., Wang, J., He, Y., 2013. BrainNet Viewer: A network 11 Figure 1: Schematic of the 12 de-noising models evaluated in the present study. For each of the 12 models indexed at left, the table details what processing procedures and confound regressors were included in the model. De-noising models were selected from the functional connectivity literature and represented a range of strategies. 12 Figure 2: Number of edges significantly related to motion after de-noising. Successful de-noising strategies reduced the relationship between connectivity and motion. The number of edges (network connections) for which this relationship persists provides evidence of a pipeline's efficacy. A, The percentage of edges significantly related to motion in a 264-node network defined by Power et al. (2011). Fewer significant edges is indicative of better performance. B, The percentage of edges significantly related to motion in a second, 333-node network defined by Gordon et al. (2016). C, Renderings of significant edges for each de-noising strategy, ranked according to efficacy. Strategies that include regression of the mean global signal are framed in blue and consistently ranked as the best performers. 13 Figure 3: Residual QC-FC correlations after de-noising. The absolute median QC-FC correlation is another measure of the relationship between connectivity and motion. A, The absolute median correlation between functional connectivity and motion in a 264- node network defined by Power et al. (2011). A lower absolute median correlation indicates better performance. B, The absolute median correlation between functional connectivity and motion in a second, 333-node network defined by Gordon et al. (2016). C, Distributions of all edgewise QC-FC correlations after each de-noising strategy, ranked according to efficacy. Results largely recapitulated those reported in Figure 2, with GSR-based approaches (blue frame) collectively exhibiting the best performance. Whereas approaches that included more regressors generally yielded a narrower distribution, those approaches that included GSR tended to shift the distribution's center toward 0. 14 Figure 4: Distance-dependence of motion artifact after de-noising. The magnitude of motion artifact varies with the Euclidean distance separating a pair of nodes, with closer nodes generally exhibiting greater impact of motion on connectivity. A, The residual distance-dependence of motion artifact in a 264-node network defined by Power et al. (2011) following confound regression. B, The residual distance-dependence of motion artifact in a second, 333-node network defined by Gordon et al. (2016). C, Density plots indicating the relationship between the Euclidean distance separating each pair of nodes (x-axis) and the QC-FC correlation of the edge connecting those nodes (y-axis). The overall trend lines for each de-noising strategy, from which distance-dependence is computed, are indicated in red. The best performing models either excised high-motion volumes (36-parameter + scrubbing) or used more localized regressors (ICA-AROMA and wmLocal). In general, approaches that made use of GSR without censoring resulted in substantial distance-dependence. This effect was driven by differential efficacy of de-noising, with effective de-noising for long range connections but not short-range connections. 15 Figure 5: Estimated loss of temporal degrees of freedom for each pipeline evaluated. Bars indicate mean number of additional regressors per confound model; error bars indicate standard deviation for models where the number of confound regressors varies by subject. High-parameter models and framewise censoring performed well overall on other benchmarks, but were also costliest in terms of temporal degrees of freedom. 16
1810.11362
1
1810
2018-10-18T21:03:48
On The Ignition, Propagation and Termination Of The Neuronal Bursting Activity During Ictogenesis In Epileptic Patients
[ "q-bio.NC" ]
Epilepsy creates a persistent increase in the probability of spontaneous seizures. An ictal episode evolves due to acute disturbance of the fine-tuned balance between excitatory vs. inhibitory inputs within a neural network in favor of excitation. The current literature that proposes the activity-dependent disinhibition as a valid mechanism of chronic epilepsy, does not provide clues on why this mechanism emerges only in epileptic patients and how the vicious circle resulting of an activity-dependent disinhibition in over-active ictogenic network would end. A new model, which presents chronic epilepsy as a disease of faulty architecture of the neural circuit, is discussed. Wherein; variable genetic or acquired predisposing factors drive abnormalities in the construction of multiple neural circuits resulting in an activity-dependent positive feedback excitatory loops which transform normal neural circuits into ictal foci. Such new mechanism, for igniting an activity-dependent unstable excitation with subsequent relatively stable disinhibition, leads to an ictal escape rhythm. The propagation of such bursting activity occurs either electrochemically via synaptic communication to remote susceptible circuits, or chemically via a trigger wave which recruits the non-connected proximal neurons. Termination occurs abruptly when the inhibitory interneurons functionally recover and reimpose their inhibitory effect on the ictogenic circuit to transform the escape rhythm into a normal, under-control output. The proposed model elucidates various enigmatic features of the disease; and illustrates both the end-result ictogenic mechanism arising from the wide variety of etiologies of human spontaneous and acquired epilepsy, and the timing of episodic transitions from normal activity to seizures.
q-bio.NC
q-bio
On The Ignition, Propagation and Termination Of The Neuronal Bursting Activity During Ictogenesis In Epileptic Patients Eslam Abbas Harmel Hospital, 987 Nile Corniche, Al Kafor, Cairo, 11559, Egypt. Abstract Epilepsy creates a persistent increase in the probability of spontaneous seizures. An ictal episode evolves due to acute disturbance of the fine-tuned balance between excitatory vs. inhibitory inputs within a neural network in favor of excitation. The current literature that proposes the activity-dependent disinhibition as a valid mechanism of chronic epilepsy, does not provide clues on why this mechanism emerges only in epileptic patients and how the vicious circle resulting of an activity- dependent disinhibition in over-active ictogenic network would end. A new model, which presents chronic epilepsy as a disease of faulty architecture of the neural circuit, is discussed. Wherein; variable genetic or acquired predisposing factors drive abnormalities in the construction of multiple neural circuits resulting in an activity-dependent positive feedback excitatory loops which transform normal neural circuits into ictal foci. Such new mechanism, for igniting an activity-dependent unstable excitation with subsequent relatively stable disinhibition, leads to an ictal escape rhythm. The propagation of such bursting activity occurs either electrochemically via synaptic communication to remote susceptible circuits, or chemically via a trigger wave which recruits the non-connected proximal neurons. Termination occurs abruptly when the inhibitory interneurons functionally recover and reimpose their inhibitory effect on the ictogenic circuit to transform the escape rhythm into a normal, under-control output. The proposed model elucidates various enigmatic features of the disease; and illustrates both the end- result ictogenic mechanism arising from the wide variety of etiologies of human spontaneous and acquired epilepsy, and the timing of episodic transitions from normal activity to seizures. Key words: Epilepsy; Ictogenesis, Excitatory Feedbacks; Competition Zone; and Escape Rhythm Introduction Brain electrical activity is non-synchronous (McPhee and Hammer, 1995) and regulated by factors within the neuron and the extraneuronal environment. Neuronal factors comprise the type, number and distribution of ion channels; changes to receptors of neurotransmitters and modulation of gene expression (Bromfield et al., 2006). The extraneuronal factors include ion concentra- tion, synaptic plasticity and regulation of transmitters' breakdown and reuptake (Blumenfeld, 2005). During seizures; a discrete group of neurons begins abnormal excessive firing in a synchronized manner (Da Silva et al., 2003)(Margineanu, 2010) which then can propagate to neighboring regions (Gotz-Trabert et al., 2008). The characteristics of said firing are high frequency bursts of action potentials and hypersyn- chronization (Fisher et al., 2005). At the level of sin- gle neurons, the ictal discharge shows a characteristic paroxysmal depolarization shift, which consists of a se- quence of sustained neuronal depolarization resulting in a burst of action potentials, a plateau-like phase associ- ated with completion of the action potential bursts, and then rapid repolarization followed by hyperpolarization (Misulis and Head, 2003). The exact mechanism by which such smooth normal brain activity is suddenly shifted to the bursting ictal firing is still unclear (Noebels et al., 2012). The ax- iomatic mechanism, which involves an acute imbalance between the excitatory discharge and the inhibitory control, needs further elaboration on the cellular and molecular levels. Mutagenic neurochemical changes in the neurotransmitters, receptors and ion channels are a valid mechanism to address severe forms of epilep- tic encephalopathies wherein the brain activity is dis- torted with no or little normal epochs. Nevertheless; neurochemical changes on such a molecular level is time-invariant and don't provide an explanation for the unpredictable episodic, and relatively rare, seizures (Schulze-Bonhage and Kuhn, 2008) that occurs in chronic epilepsy. Besides; non-mutagenic changes in the number and distribution of ion channels and neuro- transmitter receptors can not be asserted as a primary mechanism of epilepsy's pathogenesis, or a reactive remodeling as a result of repeated ictal behavior of epileptogenic foci. Models of activity-dependent disinhibition for ictogen- esis solve the timing problem; wherein disinhibition oc- curred only at the extremes of network activity (Bracci et al., 2001)(Wester and McBain, 2014). Therefore; the probability of ictal changeover would depend on the probability of that exceptional level of activity. Likely; the same range of probabilities characterizing sponta- neous seizures can agree with the odds of these levels of activity (Staley, 2015). Once a seizure is induced by an activity-dependent disinhibition, the seizure itself can continue to produce activity levels that are sufficient to suppress inhibition, providing the necessary positive feedback to sustain the seizure. Yet; Said models do not provide the fundamental mechanism by which a robust input can transform a network with apparently normal activity into an ictal focus. Moreover; they do not give clues on how the vicious positive feedback circle will end, adding more ambiguity on the poorly understood mech- anism of termination of an ictal activity. A model for the pathogenesis of chronic epilepsy shall provide insights on the most enigmatic aspects of the dis- ease like the basic fundamental mechanism that emerges from the wide varieties of epileptogenic etiologies, prop- agation, and termination of ictal behavior; besides the unpredictable episodic timing of ictal episodes. MODEL: The ictogenesis process involves two main checkpoint steps, the initiation of the burst, and the propagation of the bursting activity. The initiation step involves abrupt imbalance between excitation and inhibition within the neural environment leading to the acute transition of normal brain activity to ictal rhythm. The proposed model functionally categorizes the origin of such abrupt imbalance into three different levels of neuronal hyper- excitability with ictal discharge, with special emphasis on the second intermediate level; the chronic epilepsy. While also proposes that the propagation checkpoint de- pends on the initial conditions; wherein if a large-enough number of neurons burst synchronously, the propagation of a depolarization wave becomes mandatory. The first level is seizures in otherwise healthy brain tis- sue, which result from the simultaneous mass-excitation of large numbers of neurons by external insult or stim- uli. Electric stimulation during ECT and chemical stimulation, either during acute overdosing or sudden withdrawal of a drug, are valid examples. During ECT; electrodes deliver an electrical impulse which traverses through intermediary brain tissue to simultaneously stimulate neurons by altering their internal electrical milieu and concentration of ions (Swartz, 2014). Chem- ical stimuli may involve toxic exposure to domoic acid, which activates excitatory GluK1 glutamate receptors (Jett, 2012), or by overdoses of theophylline, which blocks the inhibitory adenosine A1 receptor (Boison, 2011). Also; abrupt withdrawal of GABAergic-acting sedative -- hypnotic drugs can cause seizure due to chronic GABA receptor downregulation as well as glutamate overactivity, which lead to neurotransmitter sensitiza- tion and neuronal hyperexcitability (Allison and Pratt, 2003). Accordingly and as a result of both electrical and chemical stimulation; a large focus of hyperexcitable neurons bursts synchronously bypassing the initiation checkpoint to induce a depolarization wave which effec- tively propagates to induce an ictal episode, or series of episodes in case of chemical insults. The third level comprises severe epileptic encephalopathies, which result from severe genetic abnormalities that compromise important inhibitory pathways. Such mu- tations are most frequently associated with continuous altering of brain functions; wherein there are no nor- mal epochs of brain activity, with subsequent frequent seizures (Allen et al., 2013)(Veeramah et al., 2013). Said seizures are usually multiform and intractable and usually accompanied by relentless cognitive, behavioral and neurological deficits (Khan and Al Baradie, 2012). The second intermediate level is chronic epilepsy, which is emphasized in this model as it is responsible for the vast majority of seizures in humans. Chronic epilepsy can be subdivided into spontaneous, and acquired due to an acute injury of the normal brain tissue as trauma, strokes and infections. Despite different pathological ori- gin; both subdivisions leads to extemporaneous activity- dependent shifts in the balance between inhibition and excitation in one or more neural circuits to increase the probability of seizures break out. This model tracks the poorly-understood initiation, propagation and termina- tion of the ictal episode in chronic epilepsy. Within a normal neural circuit; the excitatory output of principal cells is usually put under control by the in- terneurons which mainly exert an inhibitory effect on the incoming inputs; wherein the triggering of an action po- tential is determined by summation of excitatory and in- hibitory signals. A neural circuit can be analyzed accord- ing to its neuronal content, to determine the feedback, feedforward balance between the inhibitory and the ex- citatory neurons, to yield a block diagram as shown in Fig. 1; wherein (I) represents the input excitatory sig- nal, (H) represents the input inhibitory signal from the interneurons, (e) is the summation effect and represents the actual input, (P ) is the number of excitatory prin- cipal neurons receiving the input (I) within the neural circuit, (B) is the average number of inhibitory interneu- rons for each excitatory principal neuron within the neu- 2 Fig. 1. A block diagram describes the architecture of the normal neural circuit vs. the ictogenic circuit. ral circuit, and (O) represents the output of the neural circuit. The output of the circuit is: within the neural circuit in response to the input (I). O = eP and the actual input after summation is: e = I − H then, the output of the circuit can be described as: O(1 + BP ) = IP so, the ratio between the output and the input is: O I = P 1 + BP This equation represents a negative feedback system, wherein the emergent output of the circuit is lesser than the input signal by the effect of (B, the average number of inhibitory interneurons for each excitatory principal neuron within the neural circuit). Herein and as a conse- quence of all or none law; the O I ratio actually represents the firing probability of each excitatory principal neuron 3 O I < 1 Negative feedback is a widespread inhibitory tuning mechanism within neural circuits, but negative feedfor- ward usually augments a robust control as it's charac- terized by inhibition certainty and minimal time lag. The inhibitory interneurons within a neural circuit not only put the excitatory output of the principal neu- rons under check, but also assure desynchronous firing of said neurons by decreasing the firing probability of each excitatory principal neuron within the neural cir- cuit in response to a certain input. The presence of feed inhibition guarantees that the firing probability, of an excitatory neuron in response to an input signal, is usually less than one. An ictogenic focus is characterized by defective neuronal architecture which leads to priming of positive feedback excitatory pathways consequential to interlaced excita- tory interneurons or sprouting of axonal collaterals. Said defective neuronal architecture can result from the faulty neuronal arrangement due to dysgenesis guided by ge- netic predispositions, or as an end result of the compen- satory reconstruction that occurs after brain injury. The positive feedback excitatory pathways are triggered off in an activity-dependent manner. When an episodic surge of the circuit activity reaches a critical value, the positive feedback excitatory pathways are fully activated. Then; the system enters a positive feedback loop wherein the excitatory impulses reverberate to self-reinforce further excitation. The activity-dependent positive feedback loop is ex- tremely unstable, as the value (Oj) is usually equal or more than unity (Oj ≥ 1), and collapses spontaneously; wherein (j) represents the average number of excitatory feedback interconnections for each excitatory principal neuron within the neural circuit. Yet, as a consequence of impulses reverberation and self-reinforced excitation; the interneuronal inhibitory network enters a state of stupor. When the inhibitory control is defunct (H = 0), the output of the ictogenic circuit is released. So; for each excitatory principal cell within the ictogenic cir- cuit (P = 1), each input will result in a subsequent output (I = O) forming a synchronous escape rhythm; wherein all the excitatory principal neurons receiving the input (I) within the neural circuit will burst simul- taneously upon receiving the input (I). However; such self-reinforced signal magnification and the subsequent escape rhythm (ictal discharge) occurs only at the ex- tremes of the circuit activity; wherein the probability of output release would depend on the probability of an exceptional level of a vigorous activity that reaches the critical value. The bursting activity resulting from the escape rhythm can be confined due to the integral surrounding zone of inhibition, which cannot be overcome by the current level of ictal discharge, to produce a focal abnormality of the brain electrical activity; or can propagate to distort the electrical activity of the whole or a large portion of the brain. The bursting ictal discharge will not overcome the zone of inhibition as long as the number of excitatory principal neurons, receiving an input (I) within an icto- genic neural circuit, is small (P is small) leading to fail- ure of electrochemical recruitment via synaptic commu- nication; and/or failure of chemical recruitment of the non-connected proximal neurons as (r (cid:28) φr); wherein (r) is the radius of the recruited focus, and (φr) is the minimum critical focal radius from which the chemical recruitment can begin and the trigger wave can get off. The propagation of the paroxysmal depolarization shift, as a characteristic bursting activity, can occur via synap- tic transmission to the susceptible connected neurons of other remote networks; wherein the repetitive discharge from the presynaptic ictogenic neuron leads to Ca++ ac- cumulation in the presynaptic terminals enhancing neu- rotransmitters release. Besides; it can propagate to the non-connected adjacent neurons by altering the extra- cellular ionic concentrations to produce a trigger wave which gradually recruits the neurons in the proximity to initiate a bursting activity by augmenting their ex- citability. The neuronal resting membrane potential (Vresting) is mainly maintained by the role of K + and N a+ leak chan- nels (Purves et al., 2001). Due to the high concentration level of N a+ ions in the extracellular fluid and dimin- ished membrane permeability; the effect of changes of the ion concentration in the extraneuronal environment on the resting membrane potential, differs greatly be- tween K + and N a+ ions. Such effect can be represented by the following two equations: Vresting ∝ − K + Vresting ∝ + N a+ out in − K + K + out out − N a+ N a+ out in Both equations show the two different effects of fluctua- tions in the extraneuronal concentration of K + and N a+ ions on the resting membrane potential. Accordingly and regardless of the osmotic effect; while the electrochemi- cal effect of hypo/hypernatremia is negligible, neuronal resting membrane potential is very sensitive to trivial fluctuation of extraneuronal potassium concentration. The bursting activity resulting from the escape rhythm leads to a transient increase in the extraneuronal K + concentrations within the proximity of the neural net- work. Said K + fluctuations increase the excitability of adjacent neurons due to depolarizing the resting mem- brane potential. An increase in the neuronal excitability sets the affected neurons into bursting activity, leading to the propagation of the PDS through the brain tissue. As a trigger wave, each affected neuron augments further propagation, so that the propagation of PDS doesn't slow down or lose amplitude as it travels through the brain tissue. The rate of recruitment of neurons into the bursting activity, propagation of the trigger wave, can be represented by the following equation: d(N ) dt = ρn (cid:19) (cid:18) d(Vol) (cid:18) dt (cid:19) r2. d(r) dt or wherein d(N ) dt = 4π.ρn r ≥ φr wherein (N ) is the number of recruited neurons in unit of time (t), (ρn) is the neuronal density per unit of vol- ume (Vol), and (r) is the radius of the recruited focus, or the radius of the propagating trigger wave. (φr) is the minimum critical focal radius from which the chemi- cal recruitment can launch. The bursting activity of the ictogenic circuit ends abruptly when the inhibitory in- terneurons within the ictogenic focus become function- ally effective to transform the stable escape rhythm into a normal, under-control output. 4 RESULTS: Many results can be inferred from the proposed model, but hereinafter are some that elucidate various enig- matic features of chronic epilepsy. By presenting epilepsy as a disease caused by mal-architecture of the neural circuits within the brain tissue; the probabil- ity of formation of spontaneous ictogenic foci increases within the brain regions characterized by high-rate of continuous active modulation of tissue micro-structure. Additionally; the model probes into the genesis of the ictal activity in epilep- tic patients; wherein the initiation involves igniting an activity-dependent unstable positive feedback excita- tory loop which spontaneously collapses, yet leads to a synchronous escape rhythm forming a bursting activity. The action potential generated by uninhibited ig- nition tends to have echoic reverberations; so that the resulting escape rhythm shows the characteristic multi-spiking paroxysmal depolarization shift as a hall- mark of epilepsy on the cellular level. The ictal bursting activity can be either confined by an intact zone of in- hibition, which can't be overcome by the current level of ictal discharge (P is small and/or r (cid:28) φr), to pro- duce a focal abnormality of the brain electrical activ- ity; or propagating to distort the electrical activity of the whole or a large portion of the brain. The propa- gation of the bursting discharge occurs either electro- chemically via synaptic communication to remote sus- ceptible circuits, or chemically via a trigger wave which recruit the non-connected proximal neurons. Termina- tion occurs abruptly when the inhibitory interneurons functionally recover and reimpose their inhibitory effect on the ictogenic circuit to transform the escape rhythm into a normal, under-control output. Two architectural factors contribute to the aggressive- ness of an ictogenic focus; a major factor which is the average density, or average number, of the excitatory feedback pathways for each excitatory principal neuron within the neural circuit (j), and a minor factor which is the ratio between the neuronal density to the (cid:1) within the ictal focus. Addition- ally; the velocity of the chemical neuronal recruitment via a trigger wave is proportional to the neuronal den- sity of the affected brain region. synaptic density(cid:0) ρn ρs Moreover; the model shows that spontaneous epilepsy is a dynamic non-static chronic disease; wherein untreated recurrent attacks can lead to formation of neo-foci or exaggeration of the principle one via mal- architecting resulting of ectogenic excitotoxicity and subsequent faulty rewiring driven by genetic predispo- sitions. On the other hand; a rare possibility of long- term decomposition of an ictogenic focus, due to neural plasticity, can occur if consolidation with further ictal attacks is prevented. Besides; paradoxical therapeutic 5 effect of combined anti-epileptic drugs, which enhance the GABAergic hyperpolarization, can occur, due to the higher probability of main input magnification, if the synaptic plasticity between the main presynap- tic input and the postsynaptic membrane within the competition zone is anti-Hebbian. DISCUSSION: Several decades of experiments, which depend on pharmacologically-induced seizures, have established the idea that an imbalance between inhibitory and ex- citatory activity leads to ictal episodes (Scharfman, 2007). An acute imbalance between excitation and in- hibition is thus a valid ictogenic mechanism. Problems arise when such mechanism is extended to cover the chronic process of epileptogenesis, which creates a per- sistent rise in the probability of spontaneous seizures (Staley, 2015). The timing of seizures in chronic epilepsy is unpredictable and ictal episodes are relatively rare, representing much less than 1% of the total brain ac- tivity (Moran et al., 2004). Thus; in chronic epilepsy, an ictogenic mechanism shall also explain the timing of episodic transitions from normal to ictal brain activity. Additionally; another difficulty arises when applying the model of imbalanced inhibition and excitation to the eti- ology of chronic epilepsy, which does not usually suggest a chronic imbalance. In genetic predisposed epilepsy, analyses of the genetic etiology have occasionally found causal loss-of-functions mutations in inhibitory path- ways (Macdonald and Kang, 2012), but loss-of-function mutations are also found in several excitatory pathways (Frank et al., 2006)(Carvill et al., 2014), and the ma- jority of causal mutations involve genes that do not di- rectly alter the balance of inhibition and excitation (Ran et al., 2014). In the acquired epilepsies, seizures start af- ter an insult to a normal brain tissue as a result of stroke, trauma, or infection. Steady-state imbalances in excita- tion against inhibition, in established animal models of acquired epilepsy, are difficult to demonstrate. In the Pi- locarpine model; a damage to the inhibitory neurons is compensated by an increase in GABAergic synaptoge- nesis before the onset of seizures (Zhang et al., 2009). Compensatory glutamatergic synaptogenesis also hap- pens (Buckmaster, 2014), but steady-state network im- balances, between excitation and inhibition, are not ev- ident in experimental and human epilepsy (Sutula and Dudek, 2007). The current literature that proposes the activity- dependent disinhibition as a valid mechanism in chronic epilepsy, does not provide clues on why this mechanism emerges only in epileptic patients and how the vicious circle resulting of an activity-dependent disinhibition in over-active ictogenic network would end. The proposed model presents chronic epilepsy as a disease of faulty architecture of the micro-structure of the brain tissue; wherein variable genetic or acquired predisposing fac- tors drive abnormalities in the construction of multiple neural circuits resulting in an activity-dependent posi- tive feedback excitatory loops which transform normal neural circuits into ictal foci. Such new mechanism for igniting an activity-dependent unstable excitation with subsequent relatively stable disinhibition leads to an ictal escape rhythm. Additionally; Said model provides insights about the mechanism of propagation and ter- mination of such bursting activity. Besides; it illustrates both the end-result ictogenic mechanism arising from the wide variety of etiologies of human spontaneous and acquired epilepsy, and the timing of episodic transitions from normal activity to seizures. MOLECULAR MECHANISM OF IGNITION: Either genetic predispositions drives defective wiring in spontaneous epilepsy or after-insult dysgenesis in the acquired one; both lead to architectural flaws in one or more neural circuits transforming them into ictal foci; wherein positive feedback excitatory loops are formed via sprouting axonal collaterals and/or interlaced exci- tatory interneurons. Consequently; a competition zone is formed, which comprises presynaptic long-term po- tentiated main excitatory input synapse(s), presynaptic inhibitory synapse(s) (formed by feedback or feedfor- ward inhibitory interneurons), presynaptic long-term depressed feedback excitatory synapse(s) (formed by positive feedback interneurons or axonal collaterals), and a postsynaptic membrane of a principle neuron. The competition zone is not an anatomical but a functional entity; wherein a dynamic contest among synaptic plas- ticities determine the polarization of the postsynaptic membrane and further generation of an action potential. Normally; when a main excitatory input comes into the competition zone, the activity of the inhibitory feed- forward or feedback control dictates the probability of action potential generation and firing of the postsynap- tic neuron, whereas the long-term depressed excitatory feedback plays a negligible role. Spike-timing-dependent plasticity causes a difference in the homosynaptic plasticity between the presynaptic main excitatory input, which evolves a long-term po- tentiation, and the presynaptic excitatory feedback(s), which evolve long-term depression. Due to time lag; the presynaptic feedback excitatory synapse(s) are long- term depressed because they always fires immediately after the firing of the supplied postsynaptic principal neuron, hence this particular feedback excitatory stim- ulus is made weaker (Song et al., 2000)(Debanne et al., 1994). Additionally; the rise in the synaptic weight of the presynaptic main excitatory pathway dictates a reduction in the synaptic weight of the presynap- tic feedback excitatory synapse(s) to keep the average synaptic weight approximately conserved (Lynch et al., 1977)(Chistiakova et al., 2014). A strong input form the presynaptic main excitatory pathway can be cumulated via spatial or temporal summation to surpass a threshold that ignites an ictal activity. Upon receiving a robust input, the activity- dependent depressed positive feedbacks are fully acti- vated, due to Ca++ rush, and the accumulated vesicles fuse into the presynaptic membrane to release an ex- cessive amount of excitatory neurotransmitters into the competition zone. The positive feedback excitatory loops are unstable and collapse spontaneously; yet the excitatory neurotransmitters released into the competi- tion zone render the presynaptic feedback or feedforward inhibitory synapse(s) functionally obtunded. Additional molecular mechanism can also contribute to sudden in- hibitory stupor within the competition zone; excessive unstable positive feedbacks that ignite the ictal behav- ior cause the inhibitory feedbacks, not the inhibitory feedforwards, to oscillate rapidly. Intensely activated, dendritic GABAA receptors excite rather than inhibit the postsynaptic membrane (Staley et al., 1995) of the principal neuron(s) with the competition zone. The molecular mechanism of GABA-mediated mem- brane depolarization involves differential anionic con- centration shift during intense GABAA receptor activa- tion. The membrane potential is positive to the Cl− re- versal potential (Bormann et al., 1987), so that the Cl− ions flux is inward and the intracellular Cl− concentra- tion increases, but the membrane potential is negative to the HCO− 3 reversal potential (Huguenard and Alger, 1986), so that the intracellular HCO− 3 concentration decreases due to the efflux of HCO− 3 ions. The decrease in the intracellular HCO− 3 concentration, drives the dif- fusion of CO2 across the dendritic membrane, allowing continual regeneration of intracellular HCO− 3 ions by carbonic anhydrase at a rate that exceeds the removal of intracellular Cl− ions (Staley, 1994). Accordingly; The electrochemical gradient for Cl− collapses more significantly than does the HCO− 3 gradient, producing a shift in the GABAA reversal potential toward the HCO− 3 reversal potential that is sufficient to explain the depolarizing response. Due to the functional stupor of the inhibitory pathways; the presynaptic main excitatory input is released to si- multaneously activate the postsynaptic principle neu- rons forming an escape rhythm. The action potential generated by uninhibited ignition tends to have echoic reverberations along the neuronal membrane. NMDA re- ceptors are very sensitive to changes in the membrane potential; wherein extracellular M g++ and Zn++ ions bind to specific sites on the receptor, blocking the pas- sage of other cations through the ion channel. Reverber- ating depolarization of the neuronal membrane, mainly the somatic and dendritic membrane, dislodges and re- pels the M g++ and Zn++ ions from the pore, thus al- lowing a voltage-dependent flow of Ca++ and N a+ ions into the cell and K + out of the cell (Cull-Candy et al., 6 2001)(Liu and Zhang, 2000). The influx of cations cre- ates a local signal that can; in the absence of anion- dependent inhibitory control which, besides type-A K + channels, prevent reverberating depolarization; retrigger N a+ voltage-gated ion channels. Such behavior enables multiple waves of N a+ influx, creating a train of repet- itive spikes which is followed by a subsequent hyperpo- larization generated by Ca++-dependent K + channels. This response is the paroxysmal depolarization shift, which is a hallmark of epilepsy on the cellular level. MOLECULAR MECHANISM OF PROPAGATION AND TERMINATION: If the number of principal neurons of an ictogenic focus is not large enough to overcome the zone of inhibition exerted onto their axon terminals, the bursting activity would not be electrochemically transmitted via synaptic communication to the connected remote brain loci. Ad- ditionally; the synchronous activity of said small num- ber of neurons would not significantly disturb the extra- neuronal ionic environment, thus a trigger wave would not be created leading to failure of recruitment of ad- jacent non-connected neurons of nearby loci. However; when the number of principal neurons of an ictogenic fo- cus is large enough, the bursting activity makes its way through the confinement and tends to propagate. Said large number of neurons will likely overcome the zone of inhibition and transmit the ictal activity via synaptic communication to the connected remote loci. Besides; synchronous activity of a large number of princi- pal neurons within an ictal focus can potentially disturb the extraneuronal ionic environment, generating a trig- ger wave which recruits adjacent non-connected neurons of the nearby neural networks. Ictal activity, of a large number of neurons within an ictogenic focus, induces re- gional elevation of extracellular K + ions due to persis- tent efflux. As mentioned before; neuronal resting mem- brane potential is very sensitive to changes in extracel- lular K + concentration. Regional rise in the levels of ex- tracellular K + ions reduces the magnitude of potassium gradient across the cell membrane of the adjacent neu- rons; and therefore, shifts the absolute value of the rest- ing membrane potential to depolarization. The adjacent neurons become hyperexcitable and respond by produc- ing an ictal activity. The process of recruitment is self- reinforced generating a trigger wave that promulgates spatially with a defined speed. Active synaptic clear- ance brings about rebalance among the opposed neuro- transmitters within the competition zone, so that the inhibitory control reimposes its effect and decrease the probability of firing of the principle neurons within the ictal circuit demolishing the synchronous escape rhythm; consequently the ictal episode is terminated. Reactive inhibitory reimposition contributes to the altered state of consciousness after an epileptic seizure. 7 EFFECT OF ANTIEPILEPTICS ON THE COMPE- TITION ZONE: The competition zone is a functional, not an anatomical, entity; wherein the mechanism of action of antiepilep- tics can be applied to. This subsection discusses the ef- fects of antiepileptics on the competition zone because of the paradox that results from a molecular mechanism that is believed to contribute to the functional stupor of the inhibitory synapse(s), which results from the exces- sive unstable positive feedbacks that ignite an ictal be- havior. The mechanism implicates that an intensely ac- tivated, dendritic GABAA receptors excite rather than inhibit the postsynaptic membrane (Staley et al., 1995). Accordingly; drugs that enhance the effect of the neuro- transmitter GABA at the GABAA receptors, as benzo- diazepines, can effectively prevent seizures as they pre- vent the main input summation, so they prevent the ig- nition of the activity-dependent positive feedback exci- tatory loops in the first place. But; if the main input bypasses such inhibitory effect and ignites the unstable positive feedbacks, the overstimulated GABAA recep- tor, by the drug, would enhance the excitation of the postsynaptic membrane of the principal neuron(s) and functionally contribute, and further prolong, the presy- naptic inhibitory stupor. Consequently; drugs that enhance the effect of the neu- rotransmitter GABA at the GABAA receptors, as ben- zodiazepines and the like, exacerbate epileptic seizures that develop during using them as a preventive anti- convulsant. Whilst paradoxically; such medications are more effective if they are used to terminate an active attack as they boost the effect of inhibitory synapse(s) on the stable escape rhythm; which results after sponta- neous collapse of the unstable excitatory positive feed- backs. Besides, they inhibit the electrochemical propa- gation of the ictal activity by consolidating the integrity of the inhibitory zone on the synaptic terminals of the ictogenic principal neurons of the primary ictal focus. CONCLUSION: In short; chronic epilepsy is a disease of mal-architecture of the neural circuits that results from genetically- predisposed, defective wiring in the spontaneous type; or neural dysgenesis in the acquired type. Said mal- architecture leads to formation of activity-dependent excitatory positive feedbacks that can ignite an ictal activity in response to robust stimulation. Such ictal ac- tivity tends to propagate electrochemically via synaptic communication to the connected distant loci, and chem- ically via a trigger wave to the adjacent non-connected loci. The ictal activity is terminated upon reimposition of the functionally obtunded inhibitory control over the competition zone within the primary ictal focus. REFERENCES: Allen, A. S., S. F. Berkovic, P. Cossette, N. Delanty, D. Dlugos, E. E. Eichler, M. P. Epstein, T. Glauser, D. B. Goldstein, Y. Han, et al. 2013. cephalopathies. Nature, 501(7466):217. De novo mutations in epileptic en- Allison, C. and J. Pratt 2003. Neuroadaptive processes in gabaergic and glu- tamatergic systems in benzodiazepine dependence. Pharmacology & therapeutics, 98(2):171 -- 195. Blumenfeld, H. 2005. Cellular and network mechanisms of spike-wave seizures. Epilepsia, 46:21 -- 33. Boison, D. 2011. Methylxanthines, seizures, and excitotoxicity. In Methylxanthines, Pp. 251 -- 266. Springer. Bormann, J., O. P. Hamill, and B. Sakmann 1987. Mechanism of anion permeation through chan- nels gated by glycine and gamma-aminobutyric acid in mouse cultured spinal neurones. The Journal of physiology, 385(1):243 -- 286. Bracci, E., M. Vreugdenhil, S. P. Hack, and J. G. Jefferys 2001. Dynamic modulation of excitation and inhibi- tion during stimulation at gamma and beta frequen- cies in the ca1 hippocampal region. Journal of Neu- rophysiology, 85(6):2412 -- 2422. Bromfield, E. B., J. E. Cavazos, and J. I. Sirven 2006. Basic mechanisms underlying seizures and epilepsy. Buckmaster, P. S. 2014. Does mossy fiber sprouting give rise to the epileptic state? In Issues in Clinical Epileptology: A View from the Bench, Pp. 161 -- 168. Springer. Carvill, G. L., S. Weckhuysen, J. M. McMahon, C. Hart- mann, R. S. Møller, H. Hjalgrim, J. Cook, E. Ger- aghty, B. J. Oroak, S. Petrou, et al. 2014. Gabra1 and stxbp1: novel genetic causes of dravet syndrome. Neurology, 82(14):1245 -- 1253. Chistiakova, M., N. M. Bannon, M. Bazhenov, and M. Volgushev 2014. Heterosynaptic plasticity: multiple mechanisms and multiple roles. The Neuroscientist, 20(5):483 -- 498. Da Silva, F. L., W. Blanes, S. N. Kalitzin, J. Parra, P. Suffczynski, and D. N. Velis 2003. Epilepsies as dynamical diseases of brain sys- tems: basic models of the transition between normal and epileptic activity. Epilepsia, 44:72 -- 83. Debanne, D., B. Gahwiler, and S. M. Thompson 1994. Asynchronous pre-and postsynaptic activity in- duces associative long-term depression in area ca1 of the rat hippocampus in vitro. Proceedings of the Na- tional Academy of Sciences, 91(3):1148 -- 1152. Fisher, R. S., W. V. E. Boas, W. Blume, C. Elger, P. Gen- ton, P. Lee, and J. Engel Jr 2005. Epileptic seizures and epilepsy: definitions pro- posed by the international league against epilepsy (ilae) and the international bureau for epilepsy (ibe). Epilepsia, 46(4):470 -- 472. Frank, H. Y., M. Mantegazza, R. E. Westenbroek, C. A. Robbins, F. Kalume, K. A. Burton, W. J. Spain, G. S. McKnight, T. Scheuer, and W. A. Catterall 2006. Reduced sodium current in gabaergic interneu- rons in a mouse model of severe myoclonic epilepsy in infancy. Nature neuroscience, 9(9):1142. Gotz-Trabert, K., C. Hauck, K. Wagner, S. Fauser, and A. Schulze-Bonhage 2008. Spread of ictal activity in focal epilepsy. Epilep- sia, 49(9):1594 -- 1601. Huguenard, J. and B. Alger 1986. Whole-cell voltage-clamp study of the fad- ing of gaba-activated currents in acutely dissociated hippocampal neurons. Journal of Neurophysiology, 56(1):1 -- 18. Jett, D. A. 2012. Chemical toxins that cause seizures. Neurotox- icology, 33(6):1473 -- 1475. Khan, S. and R. Al Baradie 2012. Epilepsy research and treatment, 2012. Epileptic encephalopathies: an overview. Liu, Y. and J. Zhang 2000. Recent development in nmda receptors. Chinese medical journal, 113(10):948 -- 956. Lynch, G. S., T. Dunwiddie, and V. Gribkoff 1977. tic correlate of 266(5604):737. Heterosynaptic depression: a postsynap- Nature, long-term potentiation. Cull-Candy, S., S. Brickley, and M. Farrant 2001. Nmda receptor subunits: diversity, develop- ment and disease. Current opinion in neurobiology, 11(3):327 -- 335. Macdonald, R. L. and J.-Q. Kang 2012. mrna surveillance and endoplasmic reticu- lum quality control processes alter biogenesis of mu- tant gabaa receptor subunits associated with genetic epilepsies. Epilepsia, 53:59 -- 70. Staley, K. 2015. Molecular mechanisms of epilepsy. Nature neu- roscience, 18(3):367. Staley, K. J., B. L. Soldo, and W. R. Proctor 1995. Ionic mechanisms of neuronal excitation by in- hibitory gabaa receptors. Science, 269(5226):977 -- 981. Sutula, T. P. and F. E. Dudek 2007. Unmasking recurrent excitation generated by mossy fiber sprouting in the epileptic dentate gyrus: an emergent property of a complex system. Progress in brain research, 163:541 -- 563. Swartz, C. M. 2014. A mechanism of seizure induction by electricity and its clinical implications. The journal of ECT, 30(2):94 -- 97. Veeramah, K. R., L. Johnstone, T. M. Karafet, D. Wolf, R. Sprissler, J. Salogiannis, A. Barth-Maron, M. E. Greenberg, T. Stuhlmann, S. Weinert, et al. 2013. Exome sequencing reveals new causal mutations in children with epileptic encephalopathies. Epilepsia, 54(7):1270 -- 1281. Wester, J. C. and C. J. McBain 2014. Behavioral state-dependent modulation of distinct interneuron subtypes and consequences for circuit function. Current opinion in neurobiology, 29:118 -- 125. Zhang, W., R. Yamawaki, X. Wen, J. Uhl, J. Diaz, D. A. Prince, and P. S. Buckmaster 2009. Surviving hilar somatostatin interneurons en- large, sprout axons, and form new synapses with gran- ule cells in a mouse model of temporal lobe epilepsy. Journal of Neuroscience, 29(45):14247 -- 14256. Margineanu, D. G. 2010. Epileptic hypersynchrony revisited. Neurore- port, 21(15):963 -- 967. McPhee, S. J. and G. D. Hammer 1995. Pathophysiology of Disease: An Introduction to Clinical Medicine, (Lange Medical Books). USA. Misulis, K. E. and T. C. Head 2003. Essentials of clinical neurophysiology, volume 1. Garland Science. Moran, N., K. Poole, G. Bell, J. Solomon, S. Kendall, M. McCarthy, D. McCormick, L. Nashef, J. Sander, and S. Shorvon 2004. Epilepsy in the united kingdom: seizure fre- quency and severity, anti-epileptic drug utilization and impact on life in 1652 people with epilepsy. Seizure, 13(6):425 -- 433. Noebels, J., M. Avoli, M. Rogawski, R. Olsen, and A. Delgado-Escueta 2012. Jasper's Basic Mechanisms of the Epilepsies. Oxford University Press. Purves, D., G. Augustine, D. Fitzpatrick, L. Katz, A. LaMantia, J. McNamara, and S. Williams 2001. Neuroscience 2nd edition. sunderland (ma) sin- auer associates. Ran, X., J. Li, Q. Shao, H. Chen, Z. Lin, Z. S. Sun, and J. Wu 2014. Epilepsygene: a genetic resource for genes and mutations related to epilepsy. Nucleic acids research, 43(D1):D893 -- D899. Scharfman, H. E. 2007. The neurobiology of epilepsy. Current neurology and neuroscience reports, 7(4):348 -- 354. Schulze-Bonhage, A. and A. Kuhn 2008. Unpredictability of seizures and the burden of epilepsy. Seizure prediction in epilepsy: from basic mechanisms to clinical applications. Song, S., K. D. Miller, and L. F. Abbott 2000. Competitive hebbian learning through spike- timing-dependent synaptic plasticity. Nature neuro- science, 3(9):919. Staley, K. 1994. The role of an inwardly rectifying chloride con- ductance in postsynaptic inhibition. Journal of neu- rophysiology, 72(1):273 -- 284. 9
1707.08240
2
1707
2018-03-23T19:05:12
The impact of epilepsy surgery on the structural connectome and its relation to outcome
[ "q-bio.NC" ]
Temporal lobe surgical resection brings seizure remission in up to 80% of patients, with long-term complete seizure freedom in 41%. However, it is unclear how surgery impacts on the structural white matter network, and how the network changes relate to seizure outcome. We used white matter fibre tractography on preoperative diffusion MRI to generate a structural white matter network, and postoperative T1-weighted MRI to retrospectively infer the impact of surgical resection on this network. We then applied graph theory and machine learning to investigate the properties of change between the preoperative and predicted postoperative networks. Temporal lobe surgery had a modest impact on global network efficiency, despite the disruption caused. This was due to alternative shortest paths in the network leading to widespread increases in betweenness centrality post-surgery. Measurements of network change could retrospectively predict seizure outcomes with 79% accuracy and 65% specificity, which is twice as high as the empirical distribution. Fifteen connections which changed due to surgery were identified as useful for prediction of outcome, eight of which connected to the ipsilateral temporal pole. Our results suggest that the use of network change metrics may have clinical value for predicting seizure outcome. This approach could be used to prospectively predict outcomes given a suggested resection mask using preoperative data only.
q-bio.NC
q-bio
NeuroImage: Clinical 18 (2018) 202–214 Contents lists available at ScienceDirect NeuroImage: Clinical journal homepage: www.elsevier.com/locate/ynicl The impact of epilepsy surgery on the structural connectome and its relation to outcome Peter N. Taylora,b,c,⁎, Nishant Sinhaa,b, Yujiang Wanga,b,c, Sjoerd B. Vosd,e, Jane de Tisic, Anna Miserocchic, Andrew W. McEvoyc, Gavin P. Winstonc,e,1, John S. Duncanc,e,1 a Interdisciplinary Computing and Complex BioSystems Group, School of Computing Science, Newcastle University, UK b Institute of Neuroscience, Faculty of Medical Science, Newcastle University, UK c NIHR University College London Hospitals Biomedical Research Centre, UCL Institute of Neurology, Queen Square, London WC1N 3BG, UK d Translational Imaging Group, Centre for Medical Image Computing, University College London, UK e Chalfont Centre for Epilepsy, Chalfont St Peter SL9 0LR, UK T A R T I C L E I N F O A B S T R A C T Keywords: Connectome Network Temporal lobe epilepsy Surgery Machine learning Support vector machine (SVM) Background: Temporal lobe surgical resection brings seizure remission in up to 80% of patients, with long-term complete seizure freedom in 41%. However, it is unclear how surgery impacts on the structural white matter network, and how the network changes relate to seizure outcome. Methods: We used white matter fibre tractography on preoperative diffusion MRI to generate a structural white matter network, and postoperative T1-weighted MRI to retrospectively infer the impact of surgical resection on this network. We then applied graph theory and machine learning to investigate the properties of change be- tween the preoperative and predicted postoperative networks. Results: Temporal lobe surgery had a modest impact on global network efficiency, despite the disruption caused. This was due to alternative shortest paths in the network leading to widespread increases in betweenness centrality post-surgery. Measurements of network change could retrospectively predict seizure outcomes with 79% accuracy and 65% specificity, which is twice as high as the empirical distribution. Fifteen connections which changed due to surgery were identified as useful for prediction of outcome, eight of which connected to the ipsilateral temporal pole. Conclusion: Our results suggest that the use of network change metrics may have clinical value for predicting seizure outcome. This approach could be used to prospectively predict outcomes given a suggested resection mask using preoperative data only. 1. Introduction Epilepsy is a serious neurological disorder characterised by re- current unprovoked seizures affecting 1% of the population. Neurosurgical resection can bring remission in up to 80% of those with refractory focal epilepsy, with 41% remaining entirely seizure free for years (De Tisi et al., 2011). The most common type of epilepsy surgery is anterior temporal lobe resection, in which the amygdala, anterior hippocampus, and anterior temporal neocortex are removed. The commonest neurological sequelae of temporal lobe surgery are memory impairment, visual field deficits and word-finding difficulties (Jutila et al., 2002; Gooneratne et al., 2017). Recent studies have investigated surgical outcome by considering the brain as a network of connected regions. Such networks can then be subjected to quantitative analysis techniques, which measure local and global properties in networks (see Bernhardt et al. (2015) for review). Network measures that have been found to be altered in temporal lobe epilepsy (TLE) include the clustering coefficient of a region, which captures the connectedness of neighbours of a region (Bernhardt et al., 2011). Furthermore, the strength of a connection (e.g. the number of streamlines connecting two areas), or the strength of a region's con- nectivity (e.g. the number of streamlines connecting a region to all other regions) may also be altered in TLE (Besson et al., 2014a, 2014b; Taylor et al., 2015). Another measure of a network is its efficiency, which is a measure of network integration - i.e. how easy it is to travel between one region to another via direct and indirect paths, and has been shown to be altered in patients with TLE (Liu et al., 2014). Finally, regression analysis and machine learning approaches have also been ⁎ Corresponding author at: Interdisciplinary Computing and Complex BioSystems Group, School of Computing Science, Newcastle University, UK. E-mail address: [email protected] (P.N. Taylor). 1 Equal contribution. https://doi.org/10.1016/j.nicl.2018.01.028 Received 26 July 2017; Received in revised form 5 December 2017; Accepted 21 January 2018 2213-1582/ © 2018 The Author(s). Published by Elsevier Inc. This is an open access article under the CC BY license (http://creativecommons.org/licenses/BY/4.0/). P.N. Taylor et al. NeuroImage: Clinical 18 (2018) 202–214 applied to brain networks of TLE to relate them to surgical outcome (Bonilha et al., 2013; Munsell et al., 2015; Bonilha et al., 2015; Ji et al., 2015). A challenge in comparing networks across subjects is the choice of an appropriate baseline or benchmark. There are two common ap- proaches to this. One is to threshold the connectivity so that all subjects have the same number of connections. A range of thresholds are then checked, and the most significant results reported across thresholds (Zhang et al., 2011). This has the drawback of removing 'weak' but potentially important connections. A second approach is to compare the network to a random network with the same number of regions and connections. Typically this is done by either rewiring the existing net- work (Maslov and Sneppen, 2002), or by generating a new network according to predefined rules (Betzel et al., 2016; Bauer and Kaiser, 2017). Many different types of baseline networks can be used and this will therefore influence results. Recently Kuceyeski et al. (2013) introduced the network modifica- tion (NeMo) tool in the context of stroke (Kuceyeski et al., 2015a, 2015b) and multiple sclerosis (Kuceyeski et al., 2015a, 2015b). The NeMo tool is a method to enable a direct comparison between networks that undergo change. For example, in their study of stroke, the authors drew masks over stroke affected areas and overlaid this mask with data from healthy subjects. Normal connectivity from the healthy subjects, and altered connectivity (i.e. tracts which pass through the stroke mask) were calculated. This approach allowed the authors to calculate a change in connectivity metric (ChaCo), which was shown to correlate with outcomes. Since the authors use the pre-stroke network as a baseline to investigate the implied post-stroke differences, the analysis is possible without the need to generate random networks or threshold the connectivities. This obviates the need for arbitrarily chosen surro- gate networks by effectively using the patient's own network as the surrogate instead – a distinct advantage of the technique. A drawback of that study is that the tractography was derived from a cohort of healthy controls, rather than the stroke patients. Nonetheless, this framework is ideally suited to investigate changes in networks, given a well-defined alteration such as a stroke or surgery. In this study we used a ChaCo-like approach in the context of epi- lepsy surgery and addressed the following questions: What is the impact of surgery on the patient's network? How does this impact graph the- oretic properties such as region strength, network efficiency? Do these changes to patient networks correlate with surgical outcome? Although the resection masks we use in this study are derived ret- rospectively from postoperative data, our methods could in future be applied preoperatively using a mask of the intended resection. 2. Materials and methods 2.1. Patients & MRI acquisition We retrospectively studied 53 patients who underwent temporal lobe epilepsy surgery at the National Hospital for Neurology and Neurosurgery, London, United Kingdom. Full patient details can be found in Table S11, a summary is given in Table 1. Patient outcomes were defined at 12 months postoperatively, according to the ILAE classification of surgical outcomes (Wieser et al., 2001) and separated into two groups. Group 1 includes patients who were completely sei- zure free (ILAE 1), and group 2 incorporates all other possibilities (ILAE 2–6). No patient had any prior history of neurosurgery. We used a χ2 test to check for differences between outcome groups in gender, side of surgery, and evidence of hippocampal sclerosis. We applied Kruskal- Wallis test to check for differences in age between outcome groups. All patients underwent preoperative anatomical T1-weighted MRI and preoperative diffusion MRI. Postoperative T1-weighted MRI was obtained within 12 months of surgery with the exception of one patient, who was rescanned later. MRI studies were performed on a 3T GE Signa HDx scanner (General Table 1 Patient demographics and relation to outcome group. ILAE 1 ILAE 2–6 Significance N 36 (68%) 17 (32%) Male/female Left/right TLE 16/20 22/14 4/13 8/9 Age (mean, S.D./median, I.Q.R.) Hippocampal sclerosis 37, 11.6/ 39.6, 19.25 25 (69%) 41.5, 10.6/ 42.3, 10.8 10 (59%) p = 0.3597, χ2=0.839 p = 0.3353, χ2=0.923 p = 0.2374 p = 0.4460, χ2=0.5808 Electric, Waukesha, Milwaukee, WI). Standard imaging gradients with a maximum strength of 40mT m−1 and slew rate 150 T m−1 s−1 were used. All data were acquired using a body coil for transmission, and 8- channel phased array coil for reception. Standard clinical sequences were performed including a coronal T1-weighted volumetric acquisi- tion with 170 contiguous 1.1 mm-thick slices (matrix, 256 × 256; in- plane resolution, 0.9375 × 0.9375 mm). Diffusion MRI data were acquired using a cardiac-triggered single- shot spin-echo planar imaging sequence (Wheeler-Kingshott et al., 2002) with echo time = 73 ms. Sets of 60 contiguous 2.4 mm-thick axial slices were obtained covering the whole brain, with diffusion sensitizing gradients applied in each of 52 noncollinear directions (b value of 1,200mm2 s−1 [δ = 21 ms, Δ = 29 ms, using full gradient strength of 40 mT m−1]) along with 6 non-diffusion weighted scans. The gradient directions were calculated and ordered as described elsewhere (Cook et al., 2007). The field of view was 24 cm, and the acquisition matrix size was 96 × 96, zero filled to 128 × 128 during reconstruction, of 1.875 × 1.875 × 2.4 mm. The DTI acquisition time for a total of 3480 image slices was approximately 25 min (depending on subject heart rate). reconstructed voxel giving a size 2.2. Image processing 2.2.1. T1 processing Preoperative anatomical MRI was used to generate parcellated re- gions of interest (network nodes: ROIs). We used two different ap- proaches to do this, generating two different parcellation schemes. First, we used the FreeSurfer recon-all pipeline (https://surfer.nmr. mgh.harvard.edu/), which performs intensity normalization, skull stripping, subcortical volume generation, gray/white segmentation, and parcellation (Fischl, 2012). The default parcellation scheme from FreeSurfer (the Desikan-Killiany atlas (Fischl et al., 2002; Desikan et al., 2006)) contains 82 cortical ROIs and subcortical ROIs and is widely used in the literature (e.g. Munsell et al., 2015; Taylor et al., 2015). The method FreeSurfer uses to generate its ROIs uses anatomical priors based on a manually annotated dataset from healthy controls. However, this may be suboptimal in the case of disease and therefore, we use a second approach based on geodesic information flow (GIF) to generate ROIs which has the advantage of performing well even in the presence of neuropathology (Cardoso et al., 2015). Using the GIF approach, we generate 114 cortical and subcortical ROIs (Table 2). A drawback of using the GIF approach is comparison to previous studies is less straightforward since most previous work use alternative atlases. The results presented in the main manuscript use the GIF derived ROIs, while we include results using FreeSurfer derived ROIs in supplemen- tary materials to aid comparison to previous studies. 2.2.2. DWI processing Preoperative diffusion MRI data were first corrected for signal drift (Vos et al., 2016), then eddy current and movement artefacts were 203 Occipital fusiform gyrus Opercular part of the inferior frontal gyrus Orbital part of the inferior frontal gyrus Posterior cingulate gyrus Precuneus Parahippocampal gyrus Posterior insula Parietal operculum Postcentral gyrus Posterior orbital gyrus Planum polare Precentral gyrus Planum temporale Subcallosal area Superior frontal gyrus Supplementary motor cortex Supramarginal gyrus Superior occipital gyrus Superior parietal lobule Superior temporal gyrus Temporal pole OCP OFuG OpIFG OrIFG PCgG PCu PHG PIns PO PoG POrG PP PrG PT SCA SFG SMC SMG SOG SPL STG TMP gyrus Anterior insula Anterior orbital gyrus Angular gyrus Calcarine cortex Central operculum Cuneus Entorhinal area Frontal operculum Frontal pole Fusiform gyrus Gyrus rectus Inferior occipital gyrus Inferior temporal gyrus AIns AOrG AnG Calc CO Cun Ent FO FRP FuG GRe IOG ITG Lingual gyrus Lateral orbital gyrus Middle cingulate LiG LOrG MCgG gyrus Medial frontal cortex MFC MFG Middle frontal gyrus Middle occipital MOG gyrus Medial orbital gyrus Postcentral gyrus medial segment Precentral gyrus medial segment Superior frontal gyrus medial segment Middle temporal gyrus MOrG MPoG MPrG MSFG MTG P.N. Taylor et al. Table 2 Full names of the abbreviated regions of interest (ROIs). Full name Abbreviation Full name Abbreviation Anterior cingulate ACgG Occipital pole Triangular part of the inferior frontal gyrus Transverse temporal gyrus TrIFG TTG corrected using the FSL eddy_correct tool (Andersson and Sotiropoulos, 2016). The b vectors were then rotated appropriately using the 'fdt- rotate-bvecs' tool as part of FSL (Jenkinson et al., 2012; Leemans and Jones, 2009). The diffusion data were reconstructed using generalized q-sampling imaging (Yeh et al., 2010) with a diffusion sampling length ratio of 1.2. A deterministic fibre tracking algorithm (Yeh et al., 2013) was then used, allowing for crossing fibres within voxels, with seeds placed at the whole brain. Probabilistic approaches to fibre tracking (e.g. Tournier et al., 2012) have been shown to perform less well with respect to accuracy and number of false positive connections inferred.2 The choice of tractography algorithm is therefore related to whether sensitivity or specificity is more important. It was recently shown (Zalesky et al., 2016) that the introduction of false positive connections is substantially more detrimental to the calculation of graph theoretic measures such as efficiency and clustering than the introduction of false negatives. We therefore use deterministic tractography since this has been shown to have fewer false positive connections1. Default tracto- graphy parameters from the 14 February 2017 build of DSI studio software were used as follows. The angular threshold used was 60 de- grees and the step size was set to 0.9375 mm. The anisotropy threshold was determined automatically by DSI Studio. Tracks with length < 10 mm and > 300 mm were discarded. A total of 1,000,000 tracts were calculated per subject and saved in diffusion space. 2.2.3. Resection mask To draw resection masks we linearly registered the postoperative T1 2 http://www.tractometer.org/downloads/downloads/ismrm_presentation/Challenge_ ResultTractometer_updated_for_pdf_generation.pdf. 204 NeuroImage: Clinical 18 (2018) 202–214 MRI to the preoperative MRI using the FSL FLIRT tool (Jenkinson and Smith, 2001; Jenkinson et al., 2002). Resection masks were then manually drawn using the FslView software by overlaying the post- operative MRI with the preoperative MRI starting at the most anterior coronal slice, then proceeding posteriorly every three slices. Attention was given to ensure masks did not extend beyond the Sylvian fissure into inferior frontal lobe since this is unrealistic for anterior temporal resection and there was no evidence for this on post-operative MRI. Once complete, coronal slices were then joined by drawing in every sagittal slice. Masks were saved in the preoperative T1 space. Linear registration is justified in this instance because nonlinear deformation of the tissue around the resection site may lead to misleading bound- aries due to local deformations induced as part of the processing. All registrations were visually inspected in detail during the drawing of the resection mask. 2.3. Network generation To align the tracts with the ROIs we linearly registered the pre- operative T1 image to the first non-diffusion-weighted image and saved the transformation matrix using FSL FLIRT. We then multiplied every coordinate in every tract by the inverse of this transformation matrix to get the tracts in T1 space. The quality of the registration between tracts, ROIs, and the resection mask was confirmed through visual inspection for all subjects. Since networks are constructed in native space, this removes any mismatching of track types due to potential nonlinear registration issues which is advantageous compared to previous studies of network change (Kuceyeski et al., 2016). To generate preoperative connectivity matrices, we looped through all tracts and deemed two regions as connected if the two endpoints of the tract terminate in those regions. This generated a weighted connectivity (adjacency) matrix in which each entry in the matrix represents the number of streamlines connecting two regions. To generate predicted postoperative con- nectivity matrices we performed the same process as above with one exception. Any tract that had any point within the resection mask was excluded from building the matrix. The inferred postoperative network therefore always had fewer streamlines than the preoperative network and makes the assumption that remaining portions of removed tracts subserve no functionality. This reduction in streamlines alone did not explain outcome (Fig. S1). For analysis we applied a log10 transfor- mation to connection weights. Following this, right hemisphere regions (rows and columns in the matrix) were flipped in patients with a right sided resection as we investigated ipsi- and contralateral differences. All subjects had 72 regions per hemisphere. A summary of the processing pipeline to generate the GIF con- nectivity matrices is shown in Fig. 1. 2.4. Visualisation To visualise the resection mask we first linearly registered the pre- operative T1 to the MNI brain template using FSL FLIRT to generate a transformation matrix. Next, we nonlinearly registered the preoperative T1-weighted image to the MNI template using FSL FNIRT initialised with the aforementioned transformation to generate a nonlinear warp. Finally, we applied this warp to both the preoperative T1-weighted image, and mask. All images were visually inspected to check for re- gistration quality. We repeated this for all patients to generate a mask in MNI space. These masks were then loaded into MatLab, binarised (thresholded at 0.5), and summed across all subjects. This was then saved, and overlaid with the MNI brain using FSLView. Masks in MNI space were generated for visualisation purposes only and not used in the analysis. Three-dimensional projections of brain regions were produced using the centre of mass for each region and visualised using BrainNet Viewer (Xia et al., 2013). We created the scatter plots using the UniVarScatter function in Matlab (https://github.com/GRousselet/matlab_visualisation). P.N. Taylor et al. NeuroImage: Clinical 18 (2018) 202–214 Fig. 1. Processing pipeline summary for GIF matrix generation. The pipeline is applied to each subject. 2.5. Network analysis We investigated properties of the networks using well established graph theory measures that assessed the properties of a region, the properties of a connection, and the properties of the network as a whole (i.e. global). We excluded self-connections from all measures while applying the weighted and undirected versions of the functions in the Brain Connectivity Toolbox (Rubinov and Sporns, 2010). We in- vestigated measures of strength, regional communicability (sum of communicability matrix Estrada and Hatano, 2008), betweenness cen- trality, clustering, and global efficiency, since several of these have been suggested to be altered in previous studies (Bernhardt et al., 2011; Besson et al., 2014a, 2014b; Taylor et al., 2015; Liu et al., 2014). Region volume, region strength, connection strength changes were computed by dividing the post-operative network property by the pre- operative network property. For example, if node strength reduces from 10 pre-operatively to 2 in the inferred post-operative network the change measurement will be 0.2. Other measures (betweenness, clus- tering) can increase or decrease after surgery and so the difference was computed by subtracting the pre-operative value from the post-opera- tive value. 2.6. Elastic net regularisation and support vector machine classification framework We implemented a machine learning framework to investigate the association of metrics indicating changes in the region strength, region volumes, and connection strength with seizure free (ILAE 1) and non- seizure free (ILAE 2–6) outcomes. We computed the aforementioned metrics and treated them as feature vectors. Specifically, the feature space included 380, 85, and 14 features computed from proportional reduction in the connection strength, region strength, and region vo- lumes respectively for each subject. This resulted in a high-dimensional feature matrix, A ∈ ℝn×m, where m = 479 indicates the total number of features and n = 53 denotes the total number of subjects in our study. Since the number of features, m, is much larger than the number of subjects, n, we implemented a two-step classification paradigm (Munsell et al., 2015; Casanova et al., 2011). In the first step, we per- formed feature selection by implementing logistic regression with elastic net regularisation (Zou and Hastie, 2005). This gives a sparse m- dimensional weight vector, x ∈ ℝm, which minimises the following regularised logistic regression problem: min x n ∑ i 1 = w i log(1 e + i T i x a y ( ( − c )) + ) + ρ 2 x 2 2 λ + x 1 (1) where, y = (y1,y2,⋯,yn) is the n-dimensional vector representing the surgical outcomes (+1 for seizure-free outcome and −1 for nonseizure- free outcome); a Ti denotes the i − th row of feature vector A ∈ ℝn×m; wi is the weight for the i − th sample (all samples were equally weighted in our implementation); c is the scalar intercept; λ ∥ x∥1 is the l1 reg- 2 is the l2 regularisation (smooth- ularisation (sparsity) term; and 2 ness) term. We optimised the cost function in Eq. 1 by applying the LogisticR method in the sparse learning with efficient projections (SLEP) software package (Liu et al., 2009a, 2009b; Liu and Ye, 2009; Liu and Ye, 2009; Duchi et al., 2008). xρ 2 The weight vector x computed from Eq. 1 is a sparse representation of the training data set, i.e., the zero weights indicate the features which are not associated with surgical outcome, whereas the non-zero weights indicate the features associated with the surgical outcome. We derived a binary mask by setting all the non-zero weights in x to 1. Finally, we applied this mask to the high-dimensional feature space while selecting only the features associated with surgical outcome, re- sulting in a low-dimensional feature space. In the second step of our classification framework, we incorporated the aforementioned reduced feature representation in a support vector machine (SVM) classifier with a linear kernel. We designed the SVM classifier in MATLAB using the 'fitcsvm' class. Specifically, we divided our data samples into test and training sets, incorporated a leave-one- out cross-validation scheme, and computed the average accuracy, sen- sitivity, and specificity. To prevent the classifier from evaluating skewed classification performance due to the between-class im- balance-36 in the seizure-free class as opposed to 17 in the non-sei- zure-free class-we implemented a class-weighted SVM by setting all class prior probabilities as uniform similar to previous studies (e.g. Wang et al., 2012). This sets a higher misclassification penalty for the minority class as compared to the majority class, thereby preventing the classifier from learning simply from the underlying empirical data distribution (Osuna et al., 1997; Veropoulos et al., 1999; Wu and Rohini, 2004). Since our objective was to determine a minimum set of features that would most accurately separate the seizure-free outcome and the nonseizure-free outcome classes, we incorporated a leave-one-out cross- validated grid search on the l1 and l2 regularisation parameters (λ,ρ) (Munsell et al., 2015; Casanova et al., 2011). We varied the λ parameter 205 P.N. Taylor et al. NeuroImage: Clinical 18 (2018) 202–214 Fig. 2. Resection mask and region volume changes after surgery. (a) Region mask in MNI space for visualisation only (all other analyses were performed in preoperative T1 space) shows similar spatial profile between outcome groups. (b) Absolute amount of tissue resected does not relate with outcome. (c) Percentage of whole brain removed, does not relate to outcome. (d) Proportion of tissue remaining after surgery varies between subjects and regions. Here, 1 represents that the region remained intact, 0 represents that region was removed entirely. Full region names can be found in Table 2 Each dot per region is a single subject (53 per region in total). sequentially from 0.01 to 0.96 in steps of 0.05 and the ρ parameter from 0.01 to 1.96 in steps of 0.05. At each grid point, we incorporated the machine learning framework detailed above, obtained a reduced fea- ture space, and computed the performance metrics indicating the ac- curacy, sensitivity, and specificity. 3. Results Firstly, we investigate the inferred impact of surgery on network regions and secondly, on network connections. Finally, using machine learning we investigate how network change metrics relate to outcomes from surgery. 3.1. Impact of surgery on network regions 3.1.1. Region volumes All patients had similar surgical resections with the most variation between subjects on the boundary of the resection site (Fig. 2a). The absolute (Fig. 2b), and relative (Fig. 2c) volume of the resected tissue did vary between subjects, but did not explain outcome (p = 0.56 and p = 0.62 respectively – permutation test for difference between mean, 10,000 permutations used). Fig. 2d shows the proportion of tissue re- maining after resection in different brain regions. Unsurprisingly the regions which were most reduced in volume were amygdala, hippo- campus, entorhinal cortex, parahippocampal gyrus, and the temporal pole. This suggests partial disruption to multiple regions, rather than complete disruption to specific regions. Although there is variation between subjects in the proportion of each region removed, this did not predict outcome (Fig. S2). These results are reproduced for the Free- Surfer atlas in Fig. S3. 3.1.2. Region strength Each brain region is connected to other regions with streamlines. The sum of all streamlines directly connected to a region is termed the region strength - i.e. how strongly a region is directly connected to all other regions. Fig. 3 shows the impact of removing all streamlines that pass into the resected tissue by plotting the resultant change in strength. As expected, those regions that are partially removed (i.e. those showing a reduction in Fig. 2d) also have reduced strengths. The effect of surgery on other brain regions can be demonstrated with this ap- proach. For example, the ipsilateral basal forebrain, although it is ex- tratemporal and not resected, has a substantial reduction in strength for many patients. This suggests that preoperatively the basal forebrain has many of its connections with the resected tissue, rather than other areas. Several other areas beyond the resected regions also undergo alterations in strength in many subjects including the ipsilateral lateral, medial and posterior, orbital gyri among others. These results are shown in detail in Table S12(a, b), and repeated for the FreeSurfer regions in Fig. S4. It is worth noting that region strength changes shown here may not be limited to only regions partly, or wholly resected, but may also change if a portion of a tract travels through the resection site, effectively disconnecting two non-resected areas. Broadly speaking, the strengths of contralateral regions were not as substantially affected by the surgery which suggests most connections are intrahemispheric. None of the changes in node strength met sig- nificance for association with seizure-free outcome or not. 3.1.3. Region network measures The region volume change (Fig. 2) captures information solely about individual regions, while the region strength change (Fig. 3) captures information about a region and its directly connected neigh- bours. Other measures of network regions capture information re- garding aspects of their role in the wider network. The betweenness centrality of a region represents how many of the shortest paths between other regions traverse the region. Hub regions tend to occur in many of the shortest paths and have been implicated in 206 P.N. Taylor et al. NeuroImage: Clinical 18 (2018) 202–214 Fig. 3. Changes in region strength are predominantly ipsilateral. (a) Proportion of strength remaining after surgery. A value of 1 (0) indicates all streamlines connecting to that region are kept (removed). (b) An example region shows ipsilateral hippocampal connection change in the ILAE > 1 group, and the ILAE 1 group. (c) Substantially affected regions visualised in 3D MNI space. Size and colour correspond to median reduction across all patients. Colour bars in panel (c) and (a) are consistent and correspond to the proportion of strength remaining between 0 and 1. a wide range of neurological disorders (Crossley et al., 2014). If a region is surgically removed, the shortest path between regions may be via other areas. The betweenness centrality of a region can therefore either increase or decrease following surgery. Fig. 4 shows the change in betweenness centrality after surgery. Regions involved in the resection have reductions in betweenness centrality suggesting reduced integration of those areas with the wider network. However, if those regions are no longer occurring in shortest paths, the following question remains: which ones are? Fig. 4a shows that many regions have substantial increase in betweenness centrality. The redirection of shortest paths in the network following surgery ap- pears to be widely distributed among many other regions, rather than just the hubs - suggesting there are many alternative pathways (i.e. redundancy). Panel (b) in Fig. 4 shows the spatial distribution of the median change across all patients. This confirms that the increases (shown in blue) are widespread, even in contralateral regions. Because of the widespread redistribution of the shortest paths, we also evaluated global efficiency, which is the inverse of the average shortest path length of the network. Fig. 4c demonstrates the reduction in global efficiency following surgery. Given the redirection of shortest paths shown in panel (a) of the same figure, it is now easy to see why the reduction on global efficiency is only < 10% in most cases. This change does not relate to seizure outcome (p = 0.49 – permutation test for difference between mean, 10,000 permutations used). We observed similar results when using the FreeSurfer atlas (Fig. S5). We additionally investigated the change in regional clustering coefficient, a measure of the interconnectedness of neighbours of a region, following surgery. We found very subtle differences in this measure beyond the resection site following surgery (Fig. S6), which did not relate to outcome. Reductions of regional network communic- ability were most pronounced in ipsilateral amygdala, entorhinal, and temporal pole regions but also did not relate to outcome (p > 0.05, Kruskal-Wallis test, supplementary results S7). 3.2. Impact of surgery on network connections 3.2.1. Connection strength Fig. 5a shows the connections which are affected by surgery in all patients. These connections typically involve those areas which are resected. Panel (b) shows the median reduction in connectivity fol- lowing surgery. Connections in panel (a) are a subset of those in panel (b). There are many connections which are only partly (as opposed to entirely) reduced in strength. Partial reduction in strength can be considered as removing some, but not all, of the connectivity between two areas. The greatest impact was seen in connections into the ipsi- lateral posterior temporal lobe and the inferior frontal lobe, with in some patients, connections to the ipsilateral parietal lobe and con- tralateral temporal lobe and medial frontal and parietal lobes as would be expected given Fig. 3. 3.2.2. Connection network measures Similar to the regional betweenness centrality, connection (edge) betweenness centrality captures information regarding how often a connection occurs in the shortest path between other regions – i.e. how frequently that connection is traversed in the shortest paths. This value can increase or decrease following surgery, as alternative paths is taken. 207 P.N. Taylor et al. NeuroImage: Clinical 18 (2018) 202–214 Fig. 4. Widespread changes in region betweenness centrality following surgery. (a) Enhancements (positive values) and reductions (negative values) in betweenness centrality following surgery. A value of 0 indicates no change. (b) Median change of betweenness centrality shows that decreases are generally constrained to the resected regions (red) whereas increases are widespread across many other regions (blue). Colour bars in panel (b) and (a) are consistent and correspond to the change in betweenness centrality (y axis in panel a). Red (blue) indicates a decrease (increase) in betweenness centrality after surgery. (c) Global network efficiency is typically reduced by < 10% following surgery, and does not predict outcome. Fig. 6 shows the median change in betweenness centrality across all subjects. Decreases in connection betweenness centrality were broadly limited to ipsilateral temporal areas whereas the increases were more widely distributed. 3.3. Network change association with seizure outcome As an outlook to prospective applications, we used a machine learning strategy to retrospectively investigate the association between Fig. 5. Connectivity is disrupted by surgery. (a) Connections which are reduced in strength following surgery in all patients. (b) Median reduction in connectivity strength across all patients. A value of 1 (0) indicates that the median change to a connection following surgery is a 0% (100%) reduction in the number of streamlines. Only values < 1 are shown and therefore shows only connections which are reduced in strength in the majority of patients (as opposed to connections which are reduced in all patients - shown in panel a). Connections in panel (b) are a superset of those in panel (a). 208 P.N. Taylor et al. NeuroImage: Clinical 18 (2018) 202–214 Fig. 6. Widespread changes in connection betweenness centrality following surgery. Median decreases (increases) in connection betweenness centrality across all subjects shown in red (blue). network changes – region volume, region strength, and connection strength – with seizure outcome. Our machine learning approach minimises the number of features that give a high accuracy of outcome classification. As input features we incorporated all changes in volume, region strength, and connection strength and applied an elastic net algorithm to select those which were most useful. Fig. 7 shows those features selected by the algorithm. Fifteen features were selected, all of which were connections. This Fig. 7. Features derived with the elastic net algorithm. (a) Fifteen features which were found to be most informative using our machine learning pipeline. All are connections which are reduced in strength following surgery. Colour coding shows the difference in mean reduction between groups. Connections in red are reduced by surgery more in seizure-free patients than in not-seizure-free patients. (b) Correspondence with resected tissue. 209 P.N. Taylor et al. NeuroImage: Clinical 18 (2018) 202–214 suggests that incorporating knowledge of regional properties (strength and volume change) does not improve the prediction substantially. All fifteen connections in Fig. 7 are coloured according to the difference in mean value between groups. In other words, a negative value shows that the mean ILAE > 1 patient has a greater reduction in connectivity than the mean ILAE 1 patient (and vice-versa). The greatest reduction any group can have is 1 (i.e. a 100% reduction). Values are therefore bound between ± 1, and consequently not normally distributed. We therefore show the difference in mean, rather than the effect size. Eight of the features are connections involving the temporal pole. The list of selected features is given in Table S13 and are visualised in movie S14. Replication of these results using normalised data, preoperative data, or the alternative FreeSurfer parcellation did not lead to substantial im- provements in accuracy (Supplementary results S10). Those who be- came seizure free had greater reductions of connectivity into the ipsi- lateral superior temporal lobe and frontal lobe. Table 3 shows our findings including accuracy, sensitivity and specificity when using these features. Our specificity of 64.7% reflects the finding that of all patients that the model predicted would have a not-seizure-free outcome, 64.7% actually had a not-seizure-free out- come. This is substantially higher than the empirical distribution (32%). 4. Discussion In this study we have used detailed neuroanatomical information from both pre- and postoperative MRI to infer the impact of surgery on patient-specific brain networks. Our three main contributions are as follows. First, we find that the impact of surgery leads to a reduction of < 10% in efficiency in the majority of patients, due to a redirection of shortest paths. Second, there is no single feature which fully accounts for outcome with 100% accuracy. This means that there is likely not a single 'target' to aim for with surgery in this patient group that is captured by our data. Third, we have demonstrated that machine learning, in conjunction with connectivity change metrics, can produce predictions with high accuracy, and specificity which is twice as high as the empirical distribution. This means that around two thirds of the patients we predict are not seizure free, are actually not seizure free. This could help preoperatively in identifying which patients might have lower chances of becoming seizure free and put extra consideration in offering surgery or when consenting those patients. Most previous network-based studies in epilepsy typically perform group comparisons of patients relative to controls using either struc- tural networks (Kamiya et al., 2016; Taylor et al., 2015; Lemkaddem et al., 2014; de Salvo et al., 2014; Bonilha et al., 2012), functional networks (Chavez et al., 2010; Liao et al., 2010), or both (Zhang et al., 2011; Douw et al., 2015; Wirsich et al., 2016). Few have investigated alterations in global network properties with respect to surgical out- come (although see He et al., 2017 and Morgan et al., 2017 for recent examples). Bonilha and colleagues utilised preoperative structural networks inferred from diffusion MRI to investigate network properties with respect to surgical outcomes (Bonilha et al., 2013). This includes a multi-centre study using machine learning which is perhaps most si- milar to our work (Munsell et al., 2015). That study analysed pre- operative structural networks and compared several machine learning algorithms for feature selection and classification for prediction of surgical outcomes. They found that the best performing algorithm uti- lised an elastic net/support vector machine combination, and achieved a cross-site accuracy of 66%, with a same-site accuracy of 70%. We therefore chose to use broadly the same technique for classification here. Instead of investigating preoperative networks, however, we in- vestigate the difference networks - i.e. the inferred changes brought about by patient-specific surgery. These difference networks have led us to achieve a same-site accuracy of 79% in our data. We have performed a retrospective analysis of connectivity change using a mask drawn from postoperative data – i.e. the tissue that was resected. However, we envisage that in future a prospective prediction could be performed. This should be possible using the preoperative data and a mask of what the surgeon intends to resect (e.g. through the manual adaptation of an 'average mask' in the case of a standardised surgery such as anterior temporal lobe resection). Indeed, in theory a prediction is possible for any given presurgical resection plan. Fig. 8 suggests how this could be incorporated into such an evaluation, with the aim to address the question of, 'will this proposed resection render the patient seizure free?' We do not address the separate question of, 'what is the optimal resection for this patient?', which has to integrate the need to avoid damage to critical structures (e.g. Winston et al., 2012). This study has developed a pipeline for answering the first of those two questions, and retrospectively validated it with 79% accu- racy. Approaches to investigate the second of those questions have been proposed recently (Sinha et al., 2017; Goodfellow et al., 2016). 4.1. Regional change following surgery A potentially surprising result (Fig. 2), is that we found that the amount of resected tissue did not appear to predict outcome. This is in contrast to the work of (Wyler et al., 1995) and (Shamim et al., 2009) who have shown that larger resections do correlate with seizure out- comes. However, other studies have found conflicting results which accord with our own findings (Hardy et al., 2003; Keller et al., 2015; Liao et al., 2016). The variations in the literature may be due to dif- ferent populations of left/right TLE, evidence of hippocampal sclerosis, and follow-up duration. Here we have controlled for these aspects where possible with no significant differences between outcome groups (Table 1). Our finding that there is variation between subjects, in terms of proportion of regions removed, can be explained by two potential fac- tors. First, there are differences in tissue removed between subjects, and second, there are differences in regions between subjects. The latter was one of our motivations for using two different parcellations to aid comparison (the second of which is shown in supplementary materials). Table 3 Confusion matrix indicating performance of machine learning framework for predicting surgical outcomes. Actual surgical outcome Predicted outcome Seizure free = 37 Not seizure free = 16 Accuracy = 0.792 Seizure free = 36 True positive = 31 False negative = 5 (Type II Error) True Positive Rate, or Sensitivity = 0.861 False Negative Rate, or Miss rate = 0.139 210 Not seizure free = 17 False positive = 6 (Type I Error) True negative = 11 False Positive Rate or Fall out = 0.353 True Negative Rate, or Specificity = 0.647 P.N. Taylor et al. NeuroImage: Clinical 18 (2018) 202–214 Fig. 8. Potential incorporation of pipeline for prospective evaluation. Suggestion of how our approach could be used as part of the pre-surgical evaluation. From left to right: Multimodal data is acquired and evaluated at the multi-disciplinary team meeting. If surgery is recommended a proposed surgical resection mask is drawn, then fed into the connection change pipeline. The pipeline then gives a predicted post-operative network given the patient-specific mask and patient-specific presurgical DTI & T1w MRI. The pipeline uses these predicted connectivity changes, along with the pre-existing feature set to predict seizure outcome. In this example the predicted outcome would be seizure free. Furthermore, the resolution of the parcellation (i.e. the number of re- gions that the brain network is divided into) will also affect results. Here we use fairly low resolution parcellations of ≈ 100 regions in contrast to recently proposed higher resolutions (Besson et al., 2014a, 2014b; Taylor et al., 2017; Besson et al., 2017). Our reasoning for this is that in this study we compare regional properties between subjects; thus it is important that we are confident that one region in one subject is the same region in another, and can be reliably reproduced. One of the reasons that the FreeSurfer software is so popular for this is that the authors have demonstrated confidence maps at region boundaries (Fischl et al., 2004). One of the drawbacks of using such low resolution connectivity matrices however is the inability to investigate detailed within-area architecture as in (Taylor et al., 2017). Indeed, this may be an important and productive area of future research. 4.2. Network change following surgery We suggest that our finding of only a < 10% reduction in network efficiency following surgery (Fig. 4c) may explain why surgery usually leads to relatively few severe side effects. The reason for only a subtle change in efficiency is the redirection of shortest paths via other con- nections as shown in Fig. 6. We speculate that the decreases in edge betweenness correspond to those connections which undergo Wallerian degeneration. Furthermore, we speculate that connections with in- creased betweenness centrality will undergo 'strengthening', potentially reflected by increases in FA on structural imaging postoperatively. Qualitative evidence to support this has been shown by (Winston et al., 2014) who found decreases in FA in tracts involving the resected tissue and increased FA in more distant ipsilateral superior tracts within fairly short timescales of three months. Further support for this hypothesis comes from a study by (Jeong et al., 2016) who report increases in the FA of connections following frontal lobe epilepsy surgery far from the surgical site. Other studies have also made similar findings (Yogarajah et al., 2010; Pustina et al., 2014; Faber et al., 2013). Taken together, this suggests a potential rerouting of connectivity, via new shortest paths, leading to increased FA and thus avoiding major cognitive def- icits following surgery. The relationship between network efficiency, shortest path lengths, and function is not entirely clear. However, path transitivity, which is strongly related to path length has been shown to strongly correlate to functional connectivity (Goñi et al., 2014). Network efficiency itself has also been shown to relate to intelligence in functional networks in healthy controls (van den Heuvel et al., 2009). It remains to be seen how changes to network efficiency metrics following surgery relate to patient changes in brain function beyond seizure outcome. In addition to widespread network changes such as those described above, the network change pipeline also allows the inference of more localised differences. As shown in Fig. 3, we found decreases in node strength in extratemporal areas, particularly the basal forebrain in the frontal lobe. This leads us to suggest that frontal lobe function may be altered following surgery. Several studies are in agreement with this. For example, Stretton et al. (2014) found improvements in frontal lobe working memory function after surgery, while Martin et al. (2000) showed significant increases in verbal fluency following surgery. However, other studies have also found conflicting results (see Stretton and Thompson, 2012 for review). We hypothesise these improvements are related to post-surgical alterations in connectivity, and the varia- bility in the results may be down to patient variation (i.e. the large spread in Fig. 3). This effect is typically referred to as a process of functional improvement once nociferous cortex has been removed. Future studies correlating these changes in connectivity with changes in frontal lobe function will help to elucidate these relationships, poten- tially predicting not only seizure outcome, but also functional out- comes. Indeed, this has already been demonstrated in stroke where connectivity change metrics have been found to relate to functional outcomes (Kuceyeski et al., 2015a, 2015b). 4.3. Machine learning for outcome prediction Although very few have used brain network data, several studies in 211 P.N. Taylor et al. NeuroImage: Clinical 18 (2018) 202–214 recent years have begun to apply machine learning strategies for sur- gical outcome prediction. The benefits of the approach appear pro- mising. For example, (Bernhardt et al., 2015), using hippocampal measurements as features, made outcome predictions with 81% accu- racy. In another study of surgical outcomes in temporal lobe epilepsy (Memarian et al., 2015) demonstrated accuracy of 95% using various features derived from both imaging and demographic data. In addition, (Feis et al., 2013) achieved accuracy of outcome prediction of up to 96% using preoperative T1 measurements of white matter volumes. Other studies have also used machine learning algorithms, along with imaging data, to classify lateralisation of seizure onset (Focke et al., 2012; Bennett et al., 2017). Our prediction accuracy is in a broadly similar range to many of these and may be improved by including other features such as those used for generating predictive nomograms (Jehi et al., 2015). Multiple approaches exist in machine learning to attempt to im- prove the accuracy of the outcome prediction by selecting the most predictive features from a high-dimensional feature space (Tibshirani, 1996, Hoerl and Kennard, 2004, Breiman et al., 1984, Breiman et al., 1984). Here we have adopted a fairly conservative approach in which we implement elastic net regularisation (Zou and Hastie, 2005) while incorporating regularisation parameters which identify a minimum number of features important for the prediction. Even using this con- servative approach we obtain a prediction accuracy of 79.2%. It is worth noting that we impose a prior probability expectation of 50%. If we did not impose this, and instead used the empirical distribution of outcomes (68%: 32% - row 1 in Table 1), it would be easy to get ac- curacies of at least 68% by simply suggesting all patients will be seizure- free. The prediction accuracy of 79.2%, even using our highly con- servative approach, is therefore encouraging. 4.4. Strengths & weaknesses There are several strengths to this work. For example, here we have a fairly large cohort of 53 patients which were all scanned on the same scanner using the same scanning protocol. We also have good con- sistency in terms of operation type (all anterior temporal lobe resection performed by the same 2 surgeons using the same technique), and consistency in follow-up duration for outcome - 12 months. The con- sistency in follow-up duration is particularly important because out- comes can change over time (De Tisi et al., 2011). A further strength of our study, in comparison to some other network connectivity change studies, is that our analysis is performed in subject (native) space. This means nonlinear deformations of the scans to a common space, which may decrease the accuracy of the region mapping, are not necessary. Our investigation of changes in structural connectivity following surgery using a ChaCo-like approach is, to our knowledge, novel in epilepsy. Moreover, the ChaCo approach has the advantage of not requiring ar- bitrary thresholding of the network, or comparison to arbitrarily chosen random networks as a baseline – instead the patient's own pre-operative network is used. A possible weakness of this study is that we have flipped the con- nectivity matrices in right TLE patients, simplifying our analysis into ipsi- and contralateral hemispheres. Although there is precedence for this approach (Bonilha et al., 2013; Munsell et al., 2015; Ji et al., 2015) it has been suggested that left and right TLE are not simply the mirror image of each other (Besson et al., 2014a, 2014b), especially with re- spect to post-surgical cognitive deficits and resection size which is ty- pically larger in the non-dominant hemisphere. We did try to mitigate this however by ensuring the proportion of left:right patients in each outcome group are similar (p = 0.3353; Table 1). Another potential bias is the resection mask, which was manually drawn for all subjects by the same rater, with a majority subset drawn by a second rater (results S9). Variations between raters exist, but are small. The correlation of network features between raters is > 0.98, giving confidence in the reproducibility of the masks drawn here. Our choice of tractography algorithm was influenced by the intention to limit false positive connections as suggested by Zalesky et al. (2016). Despite this, there are known limitations with all approaches and this study should also be considered with the known limitations of the imaging technologies in mind. For example, the inability to resolve connection direction and potential bias for tractography to favour shorter, straighter connections (Jones, 2010). Our use of number of streamlines as a connectivity metric is based on the extensive prior literature of ChaCo analysis by Kuceyeski et al. (2013, 2015a, 2015b). However, other connection strength metrics, such as density (Taylor et al., 2015), or volume (Irimia and Van Horn, 2016) may also yield informative results. Our network change pipeline allows to infer the immediate network change brought about by surgery. It is worth highlighting that this does not necessarily represent the long-term brain network changes in the months/years that occur as a result of e.g. plasticity or degeneration. This is both a strength and a weakness of our approach and is one reason why we assess outcomes at only 12 months. The mechanisms underlying why some patients relapse to seizure recurrence only at several years after surgery (De Tisi et al., 2011) are unclear and would require longitudinal diffusion MRI data gathered over several years to investigate in this framework. The potential to use our approach pro- spectively (Fig. 8) is a distinct advantage however. 4.5. Outlook and wider implications Although we have focussed exclusively on temporal lobe epilepsy surgery here, this method could be generalized to other types of neu- rosurgery such as those for other epilepsies such as neocortical epilepsy, tumours or movement disorders. Additionally, other outcome criteria could also be predicted. For example, instead of predicting seizure outcome, functional deficits arising due to damage to known functional subnetworks (e.g. somatomotor, attention, visual) could be predicted using existing atlases (Yeo et al., 2011). Our approach would be pos- sible if preoperative imaging data is available and a proposed resection mask drawn before the surgery. Deficits for the given mask (surgery) could then be predicted for individual subjects. Additionally, multiple proposed masks (surgeries) could be tested against each other in-silico, the results for each compared, and the optimal surgery predicted. Al- gorithms for a similar approach are already being developed for in- vasive electrode recordings (Rodionov et al., 2013; Sinha et al., 2017). 4.6. Conclusion It is to be expected that outcome from surgery will depend not only on the network connectivity of the brain before surgery, but the dis- ruption to this network caused by the surgery itself. This can be mea- sured as the change in connectivity, which can be predicted for a given resection preoperatively. Our analysis shows that such connectivity change metrics can be used to retrospectively predict seizure outcome with high accuracy, and may relate to functional outcomes. Supplementary data to this article can be found online at https:// doi.org/10.1016/j.nicl.2018.01.028. Acknowledgments We are grateful to the Wolfson Foundation and the Epilepsy Society for supporting the Epilepsy Society MRI scanner. PNT was funded by Wellcome Trust (105617/Z/14/Z). PNT thanks Rob Forsyth and Helen Marley for discussion. GPW was supported by the MRC (G0802012, MR/M00841X/1). SBV is funded by the National Institute for Health Research University College London Hospitals Biomedical Research Centre (NIHR BRC UCLH/UCL High Impact Initiative). 212 P.N. Taylor et al. NeuroImage: Clinical 18 (2018) 202–214 This study was carried out in part at UCLH and the support of the NIHR-funded UCLH/UCL BRC is gratefully acknowledged. JD is sup- ported by the NIHR Senior investigator scheme. References Andersson, Jesper L.R., Sotiropoulos, Stamatios N., 2016. An integrated approach to correction for off-resonance effects and subject movement in diffusion mr imaging. NeuroImage 125, 1063–1078 (Elsevier). Bauer, Roman, Kaiser, Marcus, 2017. Nonlinear growth: an origin of hub organization in complex networks. Royal Soc. Open Sc. 4 (3), 160691 (The Royal Society). Bennett, O., Cardoso, J., Duncan, J.S., Winston, G., Ourselin, S., 2017. Learning how to See the Invisible - Using Machine Learning to Find Underlying Abnormality Patterns in Reportedly Normal MR Brain Images from Patients with Epilepsy. Proceedings. ISMRM. Bernhardt, Boris C., Chen, Zhang, He, Yong, Evans, Alan C., Bernasconi, Neda, 2011. Graph-theoretical analysis reveals disrupted small-world organization of cortical thickness correlation networks in temporal lobe epilepsy. Cereb. Cortex 21 (9), 2147–2157 (Oxford Univ Press). Bernhardt, Boris C., Hong, Seok-Jun, Bernasconi, Andrea, Bernasconi, Neda, 2015. Magnetic resonance imaging pattern learning in temporal lobe epilepsy: classification and prognostics. Ann. Neurol. 77 (3), 436–446 (Wiley Online Library). Besson, Pierre, Dinkelacker, Vera, Valabregue, Romain, Thivard, Lionel, Leclerc, Xavier, Baulac, Michel, Sammler, Daniela, et al., 2014a. Structural connectivity differences in left and right temporal lobe epilepsy. NeuroImage 100, 135–144 (Elsevier). Besson, Pierre, Lopes, Renaud, Leclerc, Xavier, Derambure, Philippe, Tyvaert, Louise, 2014b. Intra-subject reliability of the high-resolution whole-brain structural con- nectome. NeuroImage 102, 283–293 (Elsevier). Besson, Pierre, Kathleen Bandt, S., Proix, Timothée, Lagarde, Stanislas, Jirsa, Viktor K., Ranjeva, Jean-Philippe, Bartolomei, Fabrice, Guye, Maxime, 2017. Anatomic con- sistencies across epilepsies: a stereotactic-EEG informed high-resolution structural connectivity study. Brain 140 (10), 2639–2652. Betzel, Richard F., Avena-Koenigsberger, Andrea, Goñi, Joaquín, He, Ye, De Reus, Marcel A., Griffa, Alessandra, Vértes, Petra E., et al., 2016. Generative models of the human connectome. NeuroImage 124, 1054–1064 (Elsevier). Bonilha, Leonardo, Nesland, Travis, Martz, Gabriel U., Joseph, Jane E., Spampinato, Maria V., Edwards, Jonathan C., Tabesh, Ali, 2012. Medial temporal lobe epilepsy is associated with neuronal fibre loss and paradoxical increase in structural connectivity of limbic structures. J. Neurol. Neurosurg. Psychiatry 83 (9), 903–909 (BMJ Publishing Group Ltd). Bonilha, Leonardo, Helpern, Joseph A., Sainju, Rup, Nesland, Travis, Edwards, Jonathan C., Glazier, Steven S., Tabesh, Ali, 2013. Presurgical connectome and postsurgical seizure control in temporal lobe epilepsy. Neurology 81 (19), 1704–1710 (AAN Enterprises). Bonilha, Leonardo, Jensen, Jens H., Baker, Nathaniel, Breedlove, Jesse, Nesland, Travis, Lin, Jack J., Drane, Daniel L., Saindane, Amit M., Binder, Jeffrey R., Kuzniecky, Ruben I., 2015. The brain connectome as a personalized biomarker of seizure out- comes after temporal lobectomy. Neurology 84 (18), 1846–1853 (AAN Enterprises). Breiman, L., Friedman, J.H., Alshen, R.A., Stone, C.J., 1984. CART: Classification and Regression Trees. Wadsworth, Belmont, CA. Cardoso, M. Jorge, Modat, Marc, Wolz, Robin, Melbourne, Andrew, Cash, David, Rueckert, Daniel, Ourselin, Sebastien, 2015. Geodesic information flows: spatially- variant graphs and their application to segmentation and fusion. IEEE Trans. Med. Imaging 34 (9), 1976–1988 (IEEE). Casanova, Ramon, Wagner, Benjamin, Whitlow, Christopher, Williamson, Jeff, Shumaker, Sally, Maldjian, Joseph, Espeland, Mark, 2011. High dimensional classification of structural Mri Alzheimer's disease data based on large scale regularization. Front. Neuroinform. 5, 22. http://dx.doi.org/10.3389/fninf.2011.00022. Chavez, Mario, Valencia, Miguel, Navarro, Vincent, Latora, Vito, Martinerie, Jacques, 2010. Functional modularity of background activities in normal and epileptic brain networks. Phys. Rev. Lett. 104 (11), 118701 (APS). Cook, Philip A., Symms, Mark, Boulby, Philip A., Alexander, Daniel C., 2007. Optimal acquisition orders of diffusion-weighted mri measurements. J. Magn. Reson. Imaging 25 (5), 1051–1058 (Wiley Online Library). Crossley, Nicolas A., Mechelli, Andrea, Scott, Jessica, Carletti, Francesco, Fox, Peter T., McGuire, Philip, Bullmore, Edward T., 2014. The hubs of the human connectome are generally implicated in the anatomy of brain disorders. Brain 137 (8), 2382–2395 (Oxford Univ Press). De Tisi, Jane, Bell, Gail S., Peacock, Janet L., McEvoy, Andrew W., Harkness, William F.J., Sander, Josemir W., Duncan, John S., 2011. The long-term outcome of adult epilepsy surgery, patterns of seizure remission, and relapse: a cohort study. Lancet 378 (9800), 1388–1395 (Elsevier). Desikan, Rahul S., Ségonne, Florent, Fischl, Bruce, Quinn, Brian T., Dickerson, Bradford C., Blacker, Deborah, Buckner, Randy L., et al., 2006. An automated labeling system for subdividing the human cerebral cortex on Mri scans into gyral based regions of interest. NeuroImage 31 (3), 968–980 (Elsevier). Douw, Linda, DeSalvo, Matthew N., Tanaka, Naoaki, Cole, Andrew J., Liu, Hesheng, Reinsberger, Claus, Stufflebeam, Steven M., 2015. Dissociated multimodal hubs and seizures in temporal lobe epilepsy. Ann. Clin. Transl. Neurol. 2 (4), 338–352 (Wiley Online Library). Duchi, John, Shalev-Shwartz, Shai, Singer, Yoram, Chandra, Tushar, 2008. Efficient projections onto the L1-ball for learning in high dimensions. In: Proceedings of the 25th International Conference on Machine Learning. ACM, New York, NY, USA, pp. 272–279. (ICML '08). https://doi.org/10.1145/1390156.1390191. Estrada, E., Hatano, N., 2008. Communicability in complex networks. Phys. Rev. E 77 (3), 036111. Faber, Jennifer, Schoene-Bake, Jan-Christoph, Trautner, Peter, Lehe, Marec, Elger, Christian E., Weber, Bernd, 2013. Progressive fiber tract affections after temporal lobe surgery. Epilepsia 54 (4), e53-e57. Feis, Delia-Lisa, Schoene-Bake, Jan-Christoph, Elger, Christian, Wagner, Jan, Tittgemeyer, Marc, Weber, Bernd, 2013. Prediction of post-surgical seizure outcome in left mesial temporal lobe epilepsy. Neurol. Clin. 2, 903–911 (Elsevier). Fischl, Bruce, 2012. FreeSurfer. NeuroImage 62 (2), 774–781 (Elsevier). Fischl, Bruce, Salat, David H., Busa, Evelina, Albert, Marilyn, Dieterich, Megan, Haselgrove, Christian, van der Kouwe, Andre, et al., 2002. Whole brain segmentation: automated labeling of neuroanatomical structures in the human brain. Neuron 33 (3), 341–355 (Elsevier). Fischl, Bruce, van der Kouwe, André, Destrieux, Christophe, Halgren, Eric, Ségonne, Florent, Salat, David H., Busa, Evelina, et al., 2004. Automatically parcellating the human cerebral cortex. Cereb. Cortex 14 (1), 11–22 (Oxford Univ Press). Focke, Niels K., Yogarajah, Mahinda, Symms, Mark R., Gruber, Oliver, Paulus, Walter, Duncan, John S., 2012. Automated Mr image classification in temporal lobe epilepsy. NeuroImage 59 (1), 356–362 (Elsevier). Goñi, J., van den Heuvel, M.P., Avena-Koenigsberger, A., Velez de Mendizabal, N., Betzel, R.F., Griffa, A., ... Sporns, O., 2014. Resting-brain functional connectivity predicted by analytic measures of network communication. Proc. Natl. Acad. Sci. U. S. A. 111 (2), 833–838. http://dx.doi.org/10.1073/pnas.1315529111. Goodfellow, M., Rummel, C., Abela, E., Richardson, M.P., Schindler, K., Terry, J.R., 2016. Estimation of brain network ictogenicity predicts outcome from epilepsy surgery. Sci. Rep. 6. Gooneratne, I.K., Mannan, S., de Tisi, J., Gonzalez, J.C., McEvoy, A.W., Miserocchi, A., Diehl, B., Wehner, T., Bell, G.S., Sander, J.W., Duncan, J.S., 2017. Somatic compli- cations of epilepsy surgery over 25 years at a single center. Epilepsy Res. 132, 70–77. Hardy, Steven G., Miller, John W., Holmes, Mark D., Born, Donald E., Ojemann, George A., Dodrill, Carl B., Hallam, Danial K., 2003. Factors predicting outcome of surgery for intractable epilepsy with pathologically verified mesial temporal sclerosis. Epilepsia 44 (4), 565–568 (Wiley Online Library). He, X., Doucet, G.E., Pustina, D., Sperling, M.R., Sharan, A.D., Tracy, J.I., 2017. Presurgical thalamic "hubness" predicts surgical outcome in temporal lobe epilepsy. Neurology 10–1212. van den Heuvel, M.P., Stam, C.J., Kahn, R.S., Pol, H.E.H., 2009. Efficiency of functional brain networks and intellectual performance. J. Neurosci. 29 (23), 7619–7624. Hoerl, A.E., Kennard, R.W., 2004. Ridge Regression. 9780471667193 John Wiley and Sons, Inc.http://dx.doi.org/10.1002/0471667196.ess2280. Irimia, Andrei, Van Horn, John Darrell, 2016. Scale-dependent variability and quantita- tive regimes in graph-theoretic representations of human cortical networks. Brain Connect. 6 (2), 152–163. Jehi, L., Yardi, R., Chagin, K., Tassi, L., Russo, G.L., Worrell, G., ... Chauvel, P., 2015. Development and validation of nomograms to provide individualised predictions of seizure outcomes after epilepsy surgery: a retrospective analysis. Lancet Neurol. 14 (3), 283–290. Jenkinson, Mark, Smith, Stephen, 2001. A global optimisation method for robust affine registration of brain images. Med. Image Anal. 5 (2), 143–156 (Elsevier). Jenkinson, Mark, Bannister, Peter, Brady, Michael, Smith, Stephen, 2002. Improved op- timization for the robust and accurate linear registration and motion correction of brain images. NeuroImage 17 (2), 825–841 (Elsevier). Jenkinson, Mark, Beckmann, Christian F., Behrens, Timothy E.J., Woolrich, Mark W., Smith, Stephen M., 2012. Fsl. NeuroImage 62 (2), 782–790 (Elsevier). Jeong, Jeong-Won, Asano, Eishi, Juhász, Csaba, Behen, Michael E., Chugani, Harry T., 2016. Postoperative axonal changes in the contralateral hemisphere in children with medically refractory epilepsy: a longitudinal diffusion tensor imaging connectome analysis. Hum. Brain Mapp. 37 (11), 3946–3956 (Wiley Online Library). Ji, Gong-Jun, Zhang, Zhiqiang, Xu, Qiang, Wei, Wei, Wang, Jue, Wang, Zhengge, Yang, Fang, et al., 2015. Connectome reorganization associated with surgical outcome in temporal lobe epilepsy. Medicine 94 (40), e1737 (LWW). Jones, D.K., 2010. Challenges and limitations of quantifying brain connectivity in vivo with diffusion MRI. Quant. Imaging Med. Surg. 2, 341–355. Jutila, L., Immonen, A., Mervaala, E., Partanen, J., Partanen, K., Puranen, M., Kälviäinen, R., et al., 2002. Long term outcome of temporal lobe epilepsy surgery: analyses of 140 consecutive patients. J. Neurol. Neurosurg. Psychiatry 73 (5), 486–494 (BMJ Publishing Group Ltd). Kamiya, Kouhei, Amemiya, Shiori, Suzuki, Yuichi, Kunii, Naoto, Kawai, Kensuke, Akira, Kunimatsu, Nobuhito, Saito, Shigeki, Aoki, Kuni, Ohtomo, 2016. Machine learning of Dti structural brain connectomes for lateralization of temporal lobe epilepsy. Magn. Reson. Med. Sci. 15 (1), 121–129 (Japanese Society for Magnetic Resonance in Medicine). Keller, Simon S., Richardson, Mark P., Schoene-Bake, Jan-Christoph, O'muircheartaigh, Jonathan, Elkommos, Samia, Kreilkamp, Barbara, Goh, Yee Yen, Marson, Anthony G., Elger, Christian, Weber, Bernd, 2015. Thalamotemporal alteration and postoperative seizures in temporal lobe epilepsy. Ann. Neurol. 77 (5), 760–774 (Wiley Online Library). Kuceyeski, Amy, Maruta, Jun, Relkin, Norman, Raj, Ashish, 2013. The network mod- ification (nemo) tool: elucidating the effect of white matter integrity changes on cortical and subcortical structural connectivity. Brain Connect. 3 (5), 451–463 (Mary Ann Liebert, Inc. 140 Huguenot Street, 3rd Floor New Rochelle, NY 10801 USA). Kuceyeski, A.F., Vargas, W., Dayan, M., Monohan, E., Blackwell, C., Raj, A., Fujimoto, K., Gauthier, S.A., 2015a. Modeling the relationship among gray matter atrophy, ab- normalities in connecting white matter, and cognitive performance in early multiple sclerosis. Am. J. Neuroradiol. 36 (4), 702–709 (Am Soc Neuroradiology). Kuceyeski, Amy, Navi, Babak B., Kamel, Hooman, Relkin, Norman, Villanueva, Mark, Raj, 213 P.N. Taylor et al. NeuroImage: Clinical 18 (2018) 202–214 Ashish, Toglia, Joan, O'dell, Michael, Iadecola, Costantino, 2015b. Exploring the brain's structural connectome: a quantitative stroke lesion-dysfunction mapping study. Hum. Brain Mapp. 36 (6), 2147–2160 (Wiley Online Library). Kuceyeski, Amy, Navi, Babak B., Kamel, Hooman, Raj, Ashish, Relkin, Norman, Toglia, Joan, Iadecola, Costantino, O'dell, Michael, 2016. Structural connectome disruption at baseline predicts 6‐months post‐stroke outcome. Hum. Brain Mapp. 37 (7), 2587–2601. Leemans, Alexander, Jones, Derek K., 2009. The B-matrix must be rotated when cor- recting for subject motion in DTI data. Magn. Reson. Med. 61 (6), 1336–1349. Lemkaddem, Alia, Daducci, Alessandro, Kunz, Nicolas, Lazeyras, François, Seeck, Margitta, Thiran, Jean-Philippe, Vulliémoz, Serge, 2014. Connectivity and tissue microstructural alterations in right and left temporal lobe epilepsy revealed by dif- fusion spectrum imaging. Neuroimaging Clin. N. Am. 5, 349–358 (Elsevier). Liao, Wei, Zhang, Zhiqiang, Pan, Zhengyong, Mantini, Dante, Ding, Jurong, Duan, Xujun, Cheng, Luo, Lu, Guangming, Chen, Huafu, 2010. Altered functional connectivity and small-world in mesial temporal lobe epilepsy. PLoS One 5 (1), e8525 (Public Library of Science). Liao, Wei, Ji, Gong-Jun, Xu, Qiang, Wei, Wei, Wang, Jue, Wang, Zhengge, Yang, Fang, et al., 2016. Functional connectome before and following temporal lobectomy in mesial temporal lobe epilepsy. Sci. Rep. 6, 23153 (Nature Publishing Group). Liu, Jun, Ye, Jieping, 2009. Efficient Euclidean projections in linear time. In: Proceedings of the 26th Annual International Conference on Machine Learning. ACM, New York, NY, USA, pp. 657–664. ICML '09. https://doi.org/10.1145/1553374.1553459. Liu, J., Ji, S., Ye, J., 2009a. SLEP: Sparse Learning with Efficient Projections. Arizona State University. http://www.public.asu.edu/~jye02/Software/SLEP. Liu, Jun, Chen, Jianhui, Ye, Jieping, 2009b. Large-scale sparse logistic regression. In: Proceedings of the 15th Acm Sigkdd International Conference on Knowledge Discovery and Data Mining. ACM, New York, NY, USA, pp. 547–556. KDD '09. https://doi.org/10.1145/1557019.1557082. Liu, Min, Chen, Zhang, Beaulieu, Christian, Gross, Donald W., 2014. Disrupted anatomic white matter network in left mesial temporal lobe epilepsy. Epilepsia 55 (5), 674–682 (Wiley Online Library). Martin, R.C., Sawrie, S.M., Edwards, R., Roth, D.L., Faught, E., Kuzniecky, R.I., Morawetz, R.B., Gilliam, F.G., 2000. Investigation of executive function change following anterior temporal lobectomy: selective normalization of verbal fluency. Neuropsychology 14 (4), 501. Maslov, Sergei, Sneppen, Kim, 2002. Specificity and stability in topology of protein networks. Science 296 (5569), 910–913 (American Association for the Advancement of Science). Memarian, Negar, Kim, Sally, Dewar, Sandra, Engel, Jerome, Staba, Richard J., 2015. Multimodal data and machine learning for surgery outcome prediction in compli- cated cases of mesial temporal lobe epilepsy. Comput. Biol. Med. 64, 67–78 (Elsevier). Morgan, Victoria L., Englot, Dario J., Rogers, Baxter P., Landman, Bennett A., Cakir, Ahmet, Abou-Khalil, Bassel W., Anderson, Adam W., 2017. Magnetic resonance imaging connectivity for the prediction of seizure outcome in temporal lobe epilepsy. Epilepsia 58 (7), 1251–1260. Munsell, Brent C., Wee, Chong-Yaw, Keller, Simon S., Weber, Bernd, Elger, Christian, da Silva, Laura Angelica Tomaz, Nesland, Travis, Styner, Martin, Shen, Dinggang, Bonilha, Leonardo, 2015. Evaluation of machine learning algorithms for treatment outcome prediction in patients with epilepsy based on structural connectome data. NeuroImage 118, 219–230. http://dx.doi.org/10.1016/j.neuroimage.2015.06.008. Osuna, Edgar, Freund, Robert, Girosi, Federico, 1997. Support vector machines: training and applications. Libr. Technol. Rep. AIM-1602. Pustina, Dorian, Doucet, Gaelle, Evans, James, Sharan, Ashwini, Sperling, Michael, Skidmore, Christopher, Tracy, Joseph, 2014. Distinct types of white matter changes are observed after anterior temporal lobectomy in epilepsy. PLoS One 9 (8), e104211. Rodionov, Roman, Vollmar, Christian, Nowell, Mark, Miserocchi, Anna, Wehner, Tim, Micallef, Caroline, Zombori, Gergely, et al., 2013. Feasibility of multimodal 3d neuroimaging to guide implantation of intracranial eeg electrodes. Epilepsy Res. 107 (1), 91–100 (Elsevier). Rubinov, Mikail, Sporns, Olaf, 2010. Complex network measures of brain connectivity: uses and interpretations. NeuroImage 52 (3), 1059–1069 (Elsevier). de Salvo, Matthew N., Douw, Linda, Tanaka, Naoaki, Reinsberger, Claus, Stufflebeam, Steven M., 2014. Altered structural connectome in temporal lobe epilepsy. Radiology 270 (3), 842–848. http://dx.doi.org/10.1148/radiol.13131044. Shamim, Sadat, Wiggs, Edythe, Heiss, John, Sato, Susumu, Liew, Clarissa, Solomon, Jeffrey, Theodore, William H., 2009. Temporal lobectomy: resection volume, neu- ropsychological effects, and seizure outcome. Epilepsy Behav. 16 (2), 311–314 (Elsevier). Sinha, N., Dauwels, J., Kaiser, M., Cash, S.S., Brandon Westover, M., Wang, Y., Taylor, P.N., 2017. Predicting neurosurgical outcomes in focal epilepsy patients using com- putational modelling. Brain 140 (2), 319–332. Stretton, J., Thompson, P.J., 2012. Frontal lobe function in temporal lobe epilepsy. Epilepsy Res. 98 (1), 1–13. Stretton, J., Sidhu, M.K., Winston, G.P., Bartlett, P., McEvoy, A.W., Symms, M.R., Koepp, M.J., Thompson, P.J., Duncan, J.S., 2014. Working memory network plasticity after anterior temporal lobe resection: a longitudinal functional magnetic resonance imaging study. Brain 137 (5), 1439–1453. Taylor, Peter N., Han, Cheol E., Schoene-Bake, Jan-Christoph, Weber, Bernd, Kaiser, Marcus, 2015. Structural connectivity changes in temporal lobe epilepsy: spatial features contribute more than topological measures. Neuroimaging Clin. N. Am. 8, 322–328 (Elsevier). Taylor, Peter N., Wang, Yujiang, Kaiser, Marcus, 2017. Within brain area tractography suggests local modularity using high resolution connectomics. Sci. Rep. 7 (Nature Publishing Group). Tibshirani, Robert, 1996. Regression shrinkage and selection via the lasso. J. R. Stat. Soc. Ser. B Methodol. 267–288. Tournier, J., Calamante, F., Connelly, A., 2012. MRtrix: diffusion tractography in crossing fiber regions. Int. J. Imaging Syst. Technol. 22 (1), 53–66. Veropoulos, Konstantinos, Campbell, Colin, Cristianini, Nello, 1999. Controlling the sensitivity of support vector machines. In: Proceedings of the International Joint Conference on AI, pp. 55–60. Vos, Sjoerd B., Tax, Chantal M.W., Luijten, Peter R., Ourselin, Sebastien, Leemans, Alexander, Froeling, Martijn, 2016. The importance of correcting for signal drift in diffusion Mri. In: Magnetic Resonance in Medicine. Wiley Online Library. Wang, Zuoguan, Gunduz, Aysegul, Brunner, Peter, Ritaccio, Anthony L., Ji, Qiang, Schalk, Gerwin, 2012. Decoding onset and direction of movements using electrocortico- graphic (ECoG) signals in humans. Front. Neuroeng. 5. Wheeler-Kingshott, Claudia A.M., Hickman, Simon J., Parker, Geoffrey J.M., Ciccarelli, Olga, Symms, Mark R., Miller, David H., Barker, Gareth J., 2002. Investigating cer- vical spinal cord structure using axial diffusion tensor imaging. NeuroImage 16 (1), 93–102 (Elsevier). Wieser, H.G., Blume, W.T., Fish, D., Goldensohn, E., Hufnagel, A., King, D., Sperling, M.R., Lüders, H., Pedley, Timothy A., 2001. Proposal for a new classification of outcome with respect to epileptic seizures following epilepsy surgery. Epilepsia 42 (2), 282–286 (Wiley Online Library). Winston, G.P., Daga, P., Stretton, J., Modat, M., Symms, M.R., McEvoy, A.W., Ourselin, S., Duncan, J.S., 2012. Optic radiation tractography and vision in anterior temporal lobe resection. Ann. Neurol. 71 (3), 334–341. Winston, Gavin P., Stretton, Jason, Sidhu, Meneka K., Symms, Mark R., Duncan, John S., 2014. Progressive white matter changes following anterior temporal lobe resection for epilepsy. Neuroimaging Clin. N. Am. 4, 190–200 (Elsevier). Wirsich, Jonathan, Perry, Alistair, Ridley, Ben, Proix, Timothée, Golos, Mathieu, Bénar, Christian, Ranjeva, Jean-Philippe, et al., 2016. Whole-brain analytic measures of network communication reveal increased structure-function correlation in right temporal lobe epilepsy. Neuroimaging Clin. N. Am. 11, 707–718 (Elsevier). Wu, Xiaoyun, Rohini, Srihari, 2004. Incorporating prior knowledge with weighted margin support vector machines. In: Proceedings of the Tenth ACM SIGKDD International Conference on Knowledge Discovery and Data Mining. ACM. Wyler, Allen R., Hermann, Bruce P., Somes, Grant, 1995. Extent of medial temporal re- section on outcome from anterior temporal lobectomy: a randomized prospective study. Neurosurgery 37 (5), 982–991 (LWW). Xia, Mingrui, Wang, Jinhui, He, Yong, 2013. BrainNet viewer: a network visualization tool for human brain connectomics. PLoS One 8 (7), e68910 (Public Library of Science). Yeh, Fang-Cheng, Wedeen, Van Jay, Tseng, W.-Y.I., 2010. Generalized-sampling imaging. Med. Imaging IEEE Trans. 29 (9), 1626–1635 (IEEE). Yeh, Fang-Cheng, Verstynen, Timothy D., Wang, Yibao, Fernández-Miranda, Juan C., Tseng, Wen-Yih Isaac, 2013. Deterministic diffusion fiber tracking improved by quantitative anisotropy. PLoS One 8 (11), e80713 (Public Library of Science). Yeo, B.T. Thomas, Krienen, Fenna M., Sepulcre, Jorge, Sabuncu, Mert R., Lashkari, Danial, Hollinshead, Marisa, Roffman, Joshua L., et al., 2011. The organization of the human cerebral cortex estimated by intrinsic functional connectivity. J. Neurophysiol. 106 (3), 1125–1165. Yogarajah, M., Focke, N.K., Bonelli, S.B., Thompson, P., Vollmar, C., McEvoy, A.W., Alexander, D.C., Symms, M.R., Koepp, M.J., Duncan, J.S., 2010. The structural plasticity of white matter networks following anterior temporal lobe resection. Brain 133 (8), 2348–2364. Zalesky, Andrew, Fornito, Alex, Cocchi, Luca, Gollo, Leonardo L., van den Heuvel, Martijn P., Breakspear, Michael, 2016. Connectome sensitivity or specificity: which is more important? NeuroImage 142, 407–420. Zhang, Z., Liao, W., Chen, H., Mantini, D., Ding, J.R., Xu, Q., Wang, Z., et al., 2011. Altered functional–structural coupling of large-scale brain networks in idiopathic generalized epilepsy. Brain 134 (10), 2912–2928 (Oxford Univ Press). Zou, Hui, Hastie, Trevor, 2005. Regularization and variable selection via the elastic net. J. R. Stat. Soc. Ser. B (Stat Methodol.) 67 (2), 301–320. (Royal Statistical Society, Wiley). http://www.jstor.org/stable/3647580. 214
1310.5968
1
1310
2013-10-22T15:53:16
A Cell-Level Mechanism of Contrast Gain Control
[ "q-bio.NC", "physics.bio-ph" ]
The gain of neurons' responses in the auditory cortex is sensitive to contrast changes in the stimulus within a spectrotemporal range similar to their receptive fields, which can be interpreted to represent the tuning of the input to a neuron. This indicates a local mechanism of contrast gain control, which we explore with a minimal mechanistic model here. Gain control through noisy input has been observed in vitro and in a range of computational models. We investigate the behaviour of the simplest of such models to showcase gain control, a stochastic leaky integrate-and-fire (sLIF) neuron, which exhibits gain control through divisive normalisation of the input both with and without accompanying subtractive shift of the input-response curve, depending on whether input noise is proportional to or independent of its mean. To get a more direct understanding of how the input statistics change the response, we construct an analytic approximation to the firing rate of a sLIF neuron constituted of the expression for the deterministic case and a weighted average over the derived approximate steady-state distribution of conductance due to poissonian synaptic inputs. This analytic approximation qualitatively produces the same behaviour as simulations and could be extended by spectrotemporally tuned inputs to give a simple, physiological and local mechanism of contrast gain control in auditory sensing, building on recent experimental work that has hitherto only been described by phenomenological models. By comparing our weighted average firing rate curve with the commonly used sigmoidal input-response function, we demonstrate a nearly linear relationship between both the horizontal shift (or stimulus inflection point) and the inverse gain of the sigmoid and statistics derived from the sLIF model parameters, thus providing a structural constraint on the sigmoid parameter choice.
q-bio.NC
q-bio
A Cell-Level Mechanism of Contrast Gain Control Linus J. Schumacher Wolfson Centre for Mathematical Biology, Mathematical Institute, University of Oxford, Radcliffe Observatory Quarter, Woodstock Road, Oxford, OX2 6GG, United Kingdom and Computational Biology Group, Department of Computer Science, University of Oxford, Wolfson Building, Oxford, OX1 3QD, United Kingdom Geoff K. Nicholls Department of Statistics, University of Oxford, 1 South Parks Road, Oxford, OX1 3TG, United Kingdom (Dated: October 17, 2018) 3 1 0 2 t c O 2 2 ] . C N o i b - q [ 1 v 8 6 9 5 . 0 1 3 1 : v i X r a 1 Abstract The gain of neurons' responses in the auditory cortex is sensitive to contrast changes in the stimulus within a spectrotemporal range similar to their receptive fields (Rabinowitz et al., 2012), which can be interpreted to represent the tuning of the input to a neuron. This indicates a local mechanism of contrast gain control, which we explore with a minimal mechanistic model here. Gain control through noisy input has been observed in vitro (Chance et al., 2002) and in a range of computational models (Ayaz and Chance, 2009; Longtin et al., 2002). We investigate the behaviour of the simplest of such models to showcase gain control, a stochastic leaky integrate-and-fire (sLIF) neuron, which exhibits gain control through divisive normalisation of the input both with and without accompanying subtractive shift of the input-response curve, depending on whether input noise is proportional to or independent of its mean. To get a more direct understanding of how the input statistics change the response, we construct an analytic approximation to the firing rate of a sLIF neuron constituted of the expression for the deterministic case and a weighted average over the derived approximate steady-state distribution of conductance due to poissonian synaptic inputs. This analytic approximation qualitatively produces the same behaviour as simulations and could be extended by spectrotemporally tuned inputs to give a simple, physiological and local mechanism of contrast gain control in auditory sensing, building on recent experimental work that has hitherto only been described by phenomenological models (Rabinowitz et al., 2012). By comparing our weighted average firing rate curve with the commonly used sigmoidal input-response function, we demonstrate a nearly linear relationship between both the horizontal shift (or stimulus inflection point) and the inverse gain of the sigmoid and statistics derived from the sLIF model parameters, thus providing a structural constraint on the sigmoid parameter choice. I. INTRODUCTION -- CONTRAST GAIN CONTROL Animals can process sensory information over a wide range of stimulus intensity contrasts. To do so, the gain of their neural response adjusts itself to variations in spatiotemporal (in vision) or spectrotemporal contrast (in hearing). One such mechanism of gain control, divisive normal- isation, has been successfully applied in models of the visual system for a long time, but been studied much less in the auditory system. Recently it has been shown (Rabinowitz et al., 2012, 2011) that a large part of gain control of neurons in the auditory cortex can be accounted for by horizontally shifting and divisively normalising the neural response, i.e. firing rate, in proportion to the stimulus contrast (Fig. 1). The exact relationship can be described by "gain contrast ker- nels" (Rabinowitz et al., 2011). Furthermore, it was found that the gain control is most sensitive to contrast in spectrotemporal regions where the neuron's response is also high in magnitude (Ra- binowitz et al., 2012, 2011). In other words, the contrast gain control seems to operate mainly 2 within the spectrotemporal receptive field (STRF) and be insensitive to variations in stimulus con- trast outside the STRF. This phenomenological description of gain control is indicative of local computation and raises the possibility that the underlying physiological or biophysical mechanism operates on a single cell rather than on a population level. Gain control has also been observed in mathematical and computational neuron models, the simplest relevant one being the leaky integrate-and-fire (LIF) neuron with noisy synaptic input. This model springs to mind as a good candidate for a local mechanism of contrast gain control. Starting from a LIF model (as implemented in, for example, (Ayaz and Chance, 2009), but without constructing any pools of normalising or modulatory neurons) we examine gain control in response to change in input statistics, aiming to find analytical approximations to the simulation results wherever possible in order to get a better handle on the model. By identifying qualitative and quantitative similarities between the stochastic LIF model and the gain contrast kernel model, we suggest their equivalence. The stochastic LIF model is presented as a hypothesis for the mechanism that could underlie the gain contrast kernels FIG. 1. Spectrotemporal contrast kernel model by Rabinowitz et al. (2012): The stimulus spec- trogram Ltf is convolved with a linear spectrotemporal kernel kf h (equivalent to the STRF) and passed through sigmoidal response function to give the predicted firing rate yt. The horizontal shift, or stimulus inflection point, c and the inverse gain d of the sigmoid are functions of stimulus contrast σtf , and their relationship is described by a single contrast kernel κ(cd) f g ≈ kf h. Reproduced with permission. 3 II. METHODS A. Leaky integrate and fire (LIF) neurons We start with a basic neuron model that is analytically tractable: The leaky integrate-and-fire neuron model is generally regarded as the simplest description of a nerve cell that produces any realistic output. The membrane potential V follows the differential equation C dV (t) dt = I(t) − gV (t) where C is the membrane capacitance, g the membrane conductance and I(t) the input current, which could be current flowing across the membrane through ion channels or current injected in a patch-clamp experiment. When the input current is large enough, the membrane potential rises until it reaches a threshold Vth, upon which a spike is recorded and the potential is reset to Vreset = 0, representing the firing of an action potential. For a constant input current, the firing rate (number of spikes per time) is given by the expression (Koch and Segev, 1998) 0, f (I) = [tref − C g log(1 − gVth I )] I ≤ gVth I > gVth (1) −1, where tref is the refractory period, i.e., the length of time spiking is suppressed after an action potential has been fired. Note that the model as such does not exhibit a refractory period, it has to be added to achieve saturation of the firing rate. One could set tref to a physiologically realistic value on the order of 1 ms, but for the purpose of comparing to simulations we simply set it to the simulation time step ∆t. B. stochastic LIF (sLIF) neurons Having a firing rate expression for the deterministic LIF neuron, we now want to add noise to the input. An extension to the LIF model that makes it more physiological is to let the membrane conductance g vary with stochastic synaptic input, which can be excitatory or inhibitory. To compare with simulations, we adapt the implementation from Ayaz and Chance (2009) in this section, and stay close to their notation throughout this paper. Whenever a spike arrives at an excitatory (inhibitory) synapse, e.g. through a Poisson process with rate λe (λi), the conductance increases instantaneously by a fixed absolute amount ∆ge (∆gi), representing the assumption that a fixed amount of neurotransmitter is released and opens a fixed amount of postsynaptic ion 4 channels. Between spikes, the excitatory (inhibitory) part of the conductance decays to zero with time constant τge (τgi). The synaptic conductance thus evolves independently of the membrane potential and any reversal or threshold values thereof. However, the conductance terms are now no longer constant, so the differential equation describing the evolution of the membrane potential becomes C dV (t) dt = IF F (t) + gL[VL − V (t)] + ge(t)[Ee − V (t)] + gi(t)[Ei − V (t)] where Ee (Ei) is the reversal potential of the excitatory (inhibitory) ion channels, IF F is the feed- forward current and VL (= 0 earlier) is the resting membrane potential, which is also assumed to be the value of Vreset. If we shift the potential by VL and separate the voltage dependent and independent terms, we get back our original equation C dV (t) dt = I(t) − g(t)V (t) only that now we have 'effective' current and conductance: g(t) = gL + ge(t) + gi(t) I(t) = IF F (t) + ge(t)[Ee − VL] − gi(t)[VL − Ei] (2) (3) Ideally we would now directly derive a stochastic version of the deterministic firing rate equation (1), but this is intractable. Instead, we will try and approximate the distribution of ge(t) and gi(t). From this, we might be able to work out the distributions of g(t) and I(t), and so on, hopefully arriving at a distribution of f (I) -- though a simpler approach will prove far more tractable, as we will see in the rest of this section. The model complexity is deliberately kept low for the sake of tractability and suffices for the purpose of our argument. However, more complex neuron models, such as the Morris-Lecar have been shown to be equivalent to a stochastic LIF neuron (Ditlevsen and Greenwood, 2012), giving additional justification to our model choice. 1. Mean-field failure Naıvely one might try a mean-field approach and substitute the steady-state values (cid:104)ge,i(cid:105) = ∆ge,iλe,iτge,i (assuming synaptic input is Poisson distributed with rate λ) into (1), but in practice this will often give zero firing rate even when simulations show spiking. This is because even when 5 the condition (cid:104)I(cid:105) > (cid:104)g(cid:105)Vth is not satisfied, fluctations, which are not captured by the mean field approach, can drive the membrane potential above threshold. 2. Analytical approaches in the literature Decades of work have been dedicated to obtain analytical expressions for the firing rate of an sLIF neuron (Koch and Segev, 1998). Approaches include use of the Fokker-Planck equation ((Ermentrout and Terman, 2010), section 10.2, and references therein), first order corrections to the first passage time (Brunel and Sergi, 1998), the Volterra integral equation (Buonocore et al., 2010) and coupling of excitatory and inhibitory noise (Longtin et al., 2002; L´ansk´a et al., 1994; Lansk´y et al., 1995). And while these are all valuable achievements in one way or another, they typically involve integrals that can only be evaluated numerically (such as error functions (Yu and Lee, 2003; Ostojic and Brunel, 2011)), infinite sums, gamma functions or other mathematical unpleasantries. These may allow the neuroscientist to do away with running exhaustive simulations (at least for those special cases that these approaches have been tailored to), but they generally don't provide a good insight into how a change in statistics of the noisy input affect the output. For an Ornstein-Uhlenbeck (OU) neuron a linear approximation for the firing rate in terms of input mean and variance can be found (Yu and Lee, 2003; L´ansk´y and Sacerdote, 2001; Sacerdote and Giraudo, 2013). However, in the implementation of a stochastic LIF neuron here and else- where (Ayaz and Chance, 2009), the membrane Voltage is not described by an OU process. Rather, it is the membrane conductance that can be described by an OU process, as we will show in Sec- tion II C. The membrane voltage differential equation is thus coupled to the solution of this OU process. If one assumes the conductance becomes a random variable itself, we suspect one could arrive at a Chan-Karolyi-Longstaff-Sanders (CKLS), or Brennan-Schwartz (BS), process (Chan et al., 1992; Brennan and Schwartz, 1980), in which voltage fluctuations are dependent on the input statistics and the voltage itself. To our knowledge, the mathematical behaviour of such a CKLS/BS neuron has never been studied. 6 3. Parameter values used When simulating the sLIF model to test our analytical approximations, we used the following parameter values, as in (Ayaz and Chance, 2009). Parameter Description Value Units C gL VL Vth Ee Ei ∆ge ∆gi τge,i ∆t T membrane capacitance resting (or leak) conductance resting (and reset) membrane potential threshold potential for firing an action potential reversal potential of excitatory conductance reversal potential of inhibitory conductance instantaneous change in excitatory conductance instantaneous change in inhibitory conductance time constant of synaptic conductances time step 740 20 pF nS -70 mV -52 mV 0 mV -80 mV 0.16gL 0.48gL [gL] [gL] 5 ms 0.05 ms typical length of each simulation of neural activity 1 s C. Approximations of the steady-state distribution of stochastic synaptic input To understand the influence of fluctuations in synaptic input on the firing rate of our sLIF neuron, consider the distribution of conductances, excitatory or inhibitory, here generically denoted by g. In an actual implementation of the sLIF model as described in section II B, time will be discretised and the evolution of each of the excitatory and inhibitory conductance terms can be described by an AR(1)-process of the form gt = c + φgt−∆t + σξξt where c is a constant, φ is the autoregressive parameter and ξt is white noise with variance σξ. This is based on the assumption that poissonian synaptic input can be approximated as ∆g · P(λ∆t) ≈ ∆g · N (λ∆t, √ λ∆t) = c + σξξt where P and N are Poisson and normally distributed random variables, respectively (Note we use P and N to denote both the random variable and its distribution). So c = σ2 ξ = (∆g)2λ∆t and φ = e−∆t/τg ≈ 1−∆t/τg for time steps ∆t << τg. Substituting these values into the expressions for 7 the mean µ = c/(1 − φ) and variance σ2 = σ2 ξ /(1 − φ2) of an AR(1) process and expanding to first order in ∆t/τg we see that steady-state value of each of the conductance terms is approximately normally distributed with mean and variance µg = ∆gλτg σ2 g = (∆g)2λτg/2 (4) (5) One can also write the evolution of g in the continuous time limit as an Ornstein-Uhlenbeck (OU) process. Starting from our discrete time AR(1)-process gt = (1 − ∆t/τg)gt−∆t + ∆gλ∆t + ∆g √ λ∆tξt and taking the limit ∆t → 0 we get the expression dg = Φ(µg − g)dt + σW dW where Φ = 1/τg, the mean µg is as in (4) and W is a Wiener process with variance σ2 The variance around the steady-state value in an OU-process is σ2 W = ∆g2λ . W /2Φ, which gives us the g = σ2 same value for σg as in (5). We make the steady-state assumption here, but LIF neurons with dynamic inputs have been shown to have similar gain control (Ly and Doiron, 2009). D. Weighted average of deterministic LIF neuron firing rate We derived in section II C that the excitatory and inhibitory conductances are approximately normally distributed over time. By ergodicity, or by wide-sense stationarity, the distribution of a conductance value over time is equal to distribution of conductance values at a particular time point. Hence, we can estimate the firing rate of a stochastic LIF neuron by taking the average of the firing rate expression for the deterministic case (1) over the joint probability distribution of ge and gi (assumed independent), and recall that we shifted all potential terms by VL in the sLIF model construction, i.e. f (I) ≈ (cid:104)[tref − C g log(1 − g(Vth−VL) I −1(cid:105)N (µge ,σge )·N (µgi ,σgi ), I > g(Vth − VL) )] (6) where g and I are given by equations (2) and (3), respectively, and µge,i, σge,i are given by (4,5). 8 III. RESULTS A. Agreement of analytical approximation with simulations The steady state distribution of synaptic conductances under poissonian spike input is well approximated by the normal and gamma distributions derived in the previous section (Fig. 2(a)). We can also approximate the distributions of g, I and g/I well (Fig. 7(b-d)), but as discussed in section A 2 this is really only of use if we could then go on to derive the distribution of log(1 − g(Vth − VL)/I). Approximating ge and gi as either normal or gamma distributed, one can then numerically inte- grate (1) over the (separable) joint probability distribution P (ge, gi) and thus in turn approximate FIG. 2. Steady-state distributions of the excitatory (ge) and inhibitory (gi) conductances in a sLIF neuron model. Excitatory (Ie) and inhibitory (Ii) synaptic currents are approximated by a C dV dt = d− n Vth−VL normal and a gamma distribution. The model differential equation, in terms of the parameters plotted, is −1. Distribution parameters are not fitted, but derived as described in the main text (4,5). Data shown for V and the firing rate for the deterministic case is f (I) = [tref− C(Vth−VL) log(1−r(I, g))] n(g) synaptic input rates λe = λi = 1 kHz, but approximations were qualitatively similar for rates as low as 0.5 kHz, and better the higher the rates. 9 00.511.522.500.511.522.53I/nAP Ie = ge(Ee - VL)Ii = gi(VL - Ei)NormalGamma FIG. 3. Illustration of how the weighted average of LIF firing rates is calculated to approximate the sLIF firing rate. (a) Deterministic LIF firing rates (1) for a range of conductance values (ge, gi) at synaptic input rates λe = λi = 1 kHz, colour and line thickness according to the joint probability distribution of the conductance values. The sampling density has been kept low for illustrative purposes. Black curve is the weighted average (6) of the colored curves, with the relative weight of each curve given by its colour (as in (b)) and (approximately) its line thickness. Only curves with relative weight > e−2 are shown. Other parameters as in main text, section II B 3. (b) Joint probability distribution (2D Gaussian) of (ge, gi). (c,d) As (a,b), but with λe = λi = 3 kHz. the firing rate of a sLIF neuron to a given input, as illustrated in Fig. 3. The expression for the weighted average of the deterministic LIF neuron firing rate (6) agrees well with the firing rates estimated from direct simulation of the stochastic LIF equations and captures the decrease in gain when noisiness of the synaptic input is increased (Fig. 4). B. Relating parameters of the sigmoid non-linearity to those of the sLIF model Having an analytical approximation for the firing rate of a sLIF neuron (6) that agrees well with the simulations (Fig. 4) we would like to relate changes in the synaptic input statistics to the 10 11.522.5050100firing rate/HzI/nAage/nSgi/nSb 01020300204060801000.0050.010.0150.0222.533.5050100firing rate/HzI/nAcge/nSgi/nSd 0204060800501001502002500.010.020.03 FIG. 4. Firing rate of sLIF model simulations is approximated by weighted average of deter- ministic LIF firing rate expression Firing rates of the sLIF model were estimated from simulations with low (λe = λi = 1 kHz) and high (λe = λi = 3 kHz) noise balanced synaptic input (qualitatively similar fits could be produced for rates as low as 0.5 kHz). Error bars are standard deviation over 100 simulations. Simulation results are approximated by averaging the analytical firing rate expression for the deterministic LIF model over a normal or gamma distribution for the conductances. changes in the shape of the response curve. The weighted average curves resulting from (6) are similar in shape to the sigmoid non-linearities used to describe gain control in recent experimental work (Rabinowitz et al., 2012, 2011): y(z) = a + b 1 + e− z−c d (7) where a, b, c and d are parameters. With this expression, gain control was studied by fitting the parameters of the sigmoid function as functions of stimulus contrast (Rabinowitz et al., 2012, 2011) (Fig. 1). We would like to relate the parameters of this phenomenological description of gain control to parameters in our mechanistic model. Generally this is intractable, as (6) cannot be evaluated in closed form, but we can make some progress by comparing the shape of the weighted average firing rate with that of the sigmoid, and how their shapes change with respect to their parameters. In the sLIF model described here, there is no background firing, so a = 0, although this would 11 00.511.522.533.5020406080100I/nAfiring rate/Hz λ = 1 kHzλ = 3 kHznormalgamma have to be fitted to some positive value when comparing against in vivo data. Data from in vivo experiments mainly seem to lie on the left half of the fitted sigmoid function, i.e. the abscissae are smaller than the abscissa of the inflection point (Rabinowitz et al., 2012, 2011). By inspecting Fig. 3 we see that the inflection abscissa of a sigmoid (were its left half fitted to the numerical data) roughly coincides with the threshold input to get firing (the rheobase) in the deterministic case, at mean conductance values. Intuitively, this can be explained by the fact that the deterministic firing rate curve for the mean conductances has the highest weighting in the average and hence strongly determines it shape. For abscissa greater than the mean threshold input, this main curve will be linear, i.e. with no curvature, as will the resulting average, approximately. Therefore we might be tempted to set c to the threshold vale of our input I, i.e. µth = (µge + µgi + gL)(Vth − VL) (8) Similarly one can argue that the standard deviation in the input threshold will be inversely related to the gain of the firing rate curve, because it controls the horizontal spread of deterministic firing rate curves with high weighting (Fig. 3), so one might like to try and set d equal to (cid:113) σth = σ2 ge + σ2 gi(Vth − VL) (9) Similar observations about the qualitative dependence of the firing rate curve shape on input statistics have been made by Yu and Lee (2003), but they did not attempt to relate the f-I curve parameters to the input statistics in a functional form. To test these predictions, we fitted sigmoid functions to data from sLIF simulations using Matlab's fit function. The fits were generally qualitatively indistinguishable when initialised at random multiple times with c and d constrained to be of the same order of magnitude as µth and σth, though fitting would occasionally get stuck in what appeared to be local minima. For Figs 5&6 c and d were initialised at µth and σth, respectively. A sigmoid with c given by (8) is in decent agreement with the sLIF neuron model data (Fig. 5). When d is given by (9) agreement is less good but the qualitative characteristics as the curve shape and its shifting and scaling are still captured. The quality of fits is similar when both c and d are fixed and only parameter b is fitted. 12 FIG. 5. sLIF firing rates are approximated by sigmoid functions Data from simulations of the sLIF model for five balanced synaptic input rates (λ = λe = λi) were used to fit a sigmoid (7) with a = 0 and the parameters b, c, d fitted, or with either or both c and d fixed at µth (8) and σth (9), respectively. Error bars are standard deviations of 100 simulations. 13 0.811.21.41.6010203040λ = 1f/HzI/nA1.21.41.61.820102030λ = 1.75f/HzI/nA1.822.22.42.6051015λ = 2.5f/HzI/nA2.22.42.62.830510λ = 3.25f/HzI/nA2.62.833.23.402468λ = 4f/HzI/nA sLIF datab,c,d fittedb,c fitted; d = σthb,d fitted; c = µthb fitted; c = µth, d = σth FIG. 6. Relationship between fitted sigmoid parameters and statistics of sLIF model parameters Parameters b, c and d of a sigmoid (7) on data from sLIF model simulations for a range of balanced synaptic input rates from λ = 1 kHz to 4 kHz. Error bars are 95% confidence intervals on fitted parameters, as given by Matlab's fit function (curve fitting toolbox). Red diagonal crosses without error bars indicate that the confidence interval was greater than the range of abscissae. µth = (µge + µgi + gL)(Vth − VL) and σth = (Vth − VL). (cid:113) σ2 ge + σ2 gi 14 020004000600010002000300040005000µth/pAcb c d fitted200300400500600100200300400500600σth/pAdb c d fitted02000400060000200040006000800010000µth/pAcb c fitted, d fixed200300400500600100200300400500600σth/pAdb d fitted, c fixed IV. DISCUSSION We have presented here a stochastic LIF neuron as a mechanistic model for contrast gain control in neural auditory processing. In the model, subtractive and divisive normalisation of the firing rate curve arise simply from the dynamics of noisy synaptic input. This in itself is no new observation, but it has, to our knowledge, not been made use of to understand auditory contrast gain control. Only recently has the tuning of contrast sensitivity in vivo been found to be similar to the spectrotemporal receptive field (STRF) (Rabinowitz et al., 2012, 2011). This is indicative of local computation and leaves open the possibility of contrast gain control occurring intrinsically by feed-forward mechanisms only, i.e. without recurrent feedback -- which is exactly the case in the sLIF model discussed here. Thus a neuron in the auditory cortex may not compute the stimulus contrast directly, but exhibit at least part of its gain control just through its biophysics, in response to a change in contrast in the input it receives from lower levels of auditory processing. Others have drawn similar conclusions in a more general context, and with a slightly different interpretation: Yu and Lee (2003) suggest that "the change of effective stimulus threshold in various statistical stimulus environments is the key factor underlying variance or contrast gain control". Hong et al. (2008) attribute a neuron's intrinsic gain control due to the fact that "inputs with different statistics interact with the nonlinearity of the system in different ways". A. Predictions This hypothesis makes testable predictions on the relationship between mean and variance of the input to the auditory cortex -- they ought to be proportional, as in a Poisson process -- which is something that might be measurable in the inferior colliculus, which sits upstream of the cortex in the auditory pathway. In addition, effective membrane conductance and current distributions could be measured in- tracellularly and compared to our approximations (Fig. 2). If mean conductances and internal model parameter values (section II B 3) of a cell are known, one could examine their relationship to the parameters of the sigmoid nonlinearity (fitted to in vivo data) and their dependence on stim- ulus contrast, as in (8) or (9). This could in turn motivate (or criticise) the particular choice of a sigmoid function for a neuron's response non-linearity, rather than a curve of similar shape but different functional form. The best fits of sigmoids to spike rate data seem to be achieved for models were the divisive and 15 subtractive parameters, c and d in (7), have the same stimulus contrast dependence (Rabinowitz et al., 2011). This emerges naturally from a local mechanism of gain control, as the subtrac- tive and divisive parameters are determined by the same synaptic input and hence have the same spectrotemporal tuning. Based on the arguments presented in section III B, we proposed relation- ships for the sigmoid parameters c ∝ µg and d ∝ σg, which is in qualitative agreement with other work (Ayaz and Chance, 2009; Longtin et al., 2002; Ostojic and Brunel, 2011). If these relation- ships hold, the relationship between spectrotemporal contrast kernels and receptive fields ought to be κ(c) tf ≈ ktf, as Rabinowitz et al. (2011) sug- tf = κ(d) tf ≈ ktf and κ(d) tf ≈ ktf, rather than κ(c) gested. Whether this can be confirmed upon re-inspection of previous results (Rabinowitz et al., 2011) remains to be seen. B. Model criticism To achieve purely divisive normalisation without a subtractive shift of the firing rate curve, the sLIF model needs balanced excitatory and inhibitory input (Longtin et al., 2002). This might seem like physiologically unlikely fine-tuning, but balanced synaptic input has been observed experimen- tally and simple mechanisms for maintaining the balance have been proposed (Vogels et al., 2011). Furthermore, purely divisive normalisation is rarely seen in vivo (Rabinowitz et al., 2011). Hong et al. (2008) have described gain control as a diffusion of the noiseless f-I curve, which is similar in form to our weighted average (6) for normal distributions. However, Hong et al. (2008) never see a decrease in LIF firing rate at higher noise levels, which we and others (Ayaz and Chance, 2009) do observe in our LIF neurons. This disagreement might come down to the particular relationship of mean vs. variance in the input statistics, and the resulting subtractive vs. divisive scaling of the f-I curve. However, our aim was to understand and explain the functional form of gain control seen in the auditory cortex (Rabinowitz et al., 2012, 2011), which does exhibit decreased firing rate at increased noise levels. Fitting a sigmoid non-linearity with some derived, rather than fitted, parameters seems to give good agreement with model data (Fig. 5). However, the suggested correspondences c = µth (8) and d = σth (9) appear not to be exact (Fig. 6), though (nearly) linear. Thus we might have to (a) rethink the argument presented in section III B and consider the possibility that c and d are determined by the statistics of a different set of model parameters (which might well have a similar dependence on the synaptic input rates, but with different constants of proportionality) or (b) make the fitting of the sigmoid parameters more robust. 16 C. Impact beyond auditory neuroscience Beyond the application to auditory contrast gain control, we have derived an analytical approx- imation (1) to the firing rate of a stochastic LIF neuron that, if still not in closed form, is easy to evaluate (involving numerical integration over only two dimensions) and hopefully lends itself more easily to an intuitive understanding of how the input statistics affect the firing rate than other ex- pressions (Longtin et al., 2002; Ermentrout and Terman, 2010; Brunel and Sergi, 1998; Buonocore et al., 2010; L´ansk´a et al., 1994; Lansk´y et al., 1995). D. Further work Promising directions for further work exist, most notably to fit the sLIF model on stimulus data that have been filtered through estimated STRFs (as in (Rabinowitz et al., 2012, 2011)) and to refit the gain kernel model on data generated from the sLIF model. Mathematically, one could investigate whether membrane voltage distributions can be found when its evolution is described by a stochastic process with fluctuations of voltage-dependent magnitude. E. Acknowledgements LJS would like to thank Neil Rabinowitz, Ben Willmore and Jan Schnupp for their collaboration and helpful discussions. N. C. Rabinowitz, B. D. B. Willmore, J. W. H. Schnupp, and A. J. King, The Journal of neuroscience 32, 11271 (2012). F. S. Chance, L. F. Abbott, and A. D. Reyes, Neuron 35, 773 (2002). A. Ayaz and F. S. Chance, Journal of neurophysiology 101, 958 (2009). A. Longtin, B. Doiron, and A. R. Bulsara, Bio Systems 67, 147 (2002). N. C. Rabinowitz, B. D. B. Willmore, J. W. H. Schnupp, and A. J. King, Neuron 70, 1178 (2011). C. Koch and I. Segev, Computational neuroscience (MIT Press, 1998). S. Ditlevsen and P. E. Greenwood, Journal of mathematical biology (2012), 10.1007/s00285-012-0552-7. B. Ermentrout and D. H. Terman, Mathematical Foundations of Neuroscience (Springer, 2010). N. Brunel and S. Sergi, Journal of theoretical biology 195, 87 (1998). A. Buonocore, L. Caputo, E. Pirozzi, and L. M. Ricciardi, Neural computation 22, 2558 (2010). V. L´ansk´a, P. L´ansk´y, and C. E. Smith, Journal of theoretical biology 166, 393 (1994). 17 P. Lansk´y, L. Sacerdote, and F. Tomassetti, Biological cybernetics 465, 457 (1995). Y. Yu and T. Lee, Physical Review E 68, 011901 (2003). S. Ostojic and N. Brunel, PLoS computational biology 7, e1001056 (2011). P. L´ansk´y and L. Sacerdote, Physics Letters A 285, 132 (2001). L. Sacerdote and M. T. Giraudo, Stochastic Biomathematical Models Lecture Notes in Mathematics, 2058 (2013), 10.1007/978-3-642-32157-3. K. C. Chan, G. A. Karolyi, F. A. Longstaff, and A. B. Sanders, The Journal of Finance XLVII (1992). M. Brennan and E. Schwartz, Journal of Financial and Quantitative Analysis XV (1980). C. Ly and B. Doiron, PLoS computational biology 5, e1000365 (2009). S. Hong, B. N. Lundstrom, and A. L. Fairhall, PLoS computational biology 4, e1000119 (2008). T. P. Vogels, H. Sprekeler, F. Zenke, C. Clopath, and W. Gerstner, Science 334, 1569 (2011). C. A. Coelho, Journal of Multivariate Analysis 64, 86 (1998). D. V. Hinkley, Biometrika 56, 635 (1969). R. C. Geary, Journal of the Royal Statistical Society 93, 442 (1930). Appendix A: Other approximations for the conductance distribution and beyond 1. Gamma distribution As seen in the results section (Fig. 2), the steady state distribution of the conductances seems to be captured even better by a gamma distribution with shape k and scale θ, especially with respect to its skew. Matching the mean (kθ) and variance (kθ2) of the gamma distribution to that of the normal approximation (4,5), we require the shape parameter to be k = 2λτg and the scale θ = ∆g/2. It is tempting to try to arrive at these parameters by deriving the gamma distribution as the probability distribution of the sum of k exponentially distributed random variables of mean θ, or as the distribution of waiting times until 'death' of random variables with poissonian arrival. However, the analogy does not seem to work, the parameters would be out by factors of 2 or 1/2. For this reason approximating the steady-state conductances as gamma distributed it less appealing. To arrive at an expression for the distribution of firing rates we would also like to know, as a second step, the distribution of sums and differences of the conductances (as these appear as g and I in the firing rate expression (1)), but there seems to be no established result for the difference of two gamma distributed random variables, and the sum for different shape parameters is an unwieldy expression (Coelho, 1998). 18 2. Further distributions Approximating the distribution of ge and gi as normal, we can work out the distribution of g and I, which will also be normal and (nearly) independent. The next term in the firing rate expression (1) whose distribution we would like to know is g(t)(Vth − VL)/I(t). This is approximated by a Gaussian ratio distribution (Hinkley, 1969), which can be made me made more Gaussian by a Geary-Hinkley-transformation (Hinkley, 1969; Geary, 1930) of the random variable as long as I(t) is unlikely to be negative. As seen in Fig. 7, this all works well, but going further is difficult. Working out the distribution of log(1− g(t)(Vth− VL)/I(t)) while only considering values for which g(t)(Vth − VL)/I(t) < 1 is tricky and goes beyond the scope of this paper. Instead, we turn to a more pragmatic approach in the next section. 19 FIG. 7. Steady-state distributions of terms involving the excitatory (ge) and inhibitory (gi) plotted, is C dV r(I, g))] n(g) dt = d− n Vth−VL V and the firing rate for the deterministic case is f (I) = [tref− C(Vth−VL) conductances in a sLIF neuron model. The model differential equation, in terms of the parameters log(1− −1. Distribution parameters are not fitted, but derived as described in the main text (4,5). Data shown for synaptic input rates λe = λi = 1 kHz, but approximations were qualitatively similar for rates as low as 0.5 kHz, and better the higher the rates. (a) Effective current d = IF F + Ie − Ii and effective conductance (multiplied by a voltage term) n = g(Vth − VL) are approximately normally distributed, with µd ∝ µIe − µIi, σ2 (b) The distribution of the ratio r of the terms in (b) approximately follows the ratio distribution (Hinkley, 1969). Range shown is limited to and µn ∝ µge + µgi , σ2 n = σ2 ge + σ2 gi d = σ2 Ie + σ2 Ii f (I) > 0. (c) The Geary-Hinkley-transform (Hinkley, 1969; Geary, 1930) makes r approximately normally distributed with zero mean and unit variance. Maximum abscissa as in (c). 20 0.511.522.5300.511.5I/nAPa ndN. diff.N. sum00.5100.050.1ratioPb r = n/dRatio distrib.-2-1000.010.020.03Pxc GH(r)N(0,1)
1611.04794
3
1611
2017-09-01T13:57:18
Exploring the reproducibility of functional connectivity alterations in Parkinson's Disease
[ "q-bio.NC" ]
Since anatomic MRI is presently not able to directly discern neuronal loss in Parkinson's Disease (PD), studying the associated functional connectivity (FC) changes seems a promising approach toward developing non-invasive and non-radioactive neuroimaging markers for this disease. While several groups have reported such FC changes in PD, there are also significant discrepancies between studies. Investigating the reproducibility of PD-related FC changes on independent datasets is therefore of crucial importance. We acquired resting-state fMRI scans for 43 subjects (27 patients , 16 controls) and compared the observed FC changes with those obtained in 2 independent datasets, one made available by the PPMI consortium and a second one by the group of Tao Wu. Unfortunately, PD-related functional connectivity changes turned out to be non-reproducible across datasets. This could be due to disease heterogeneity, but also to technical differences. To distinguish between the two, we devised a method to directly check for disease heterogeneity using random splits of a single dataset. Since we still observe non-reproducibility in a large fraction of random splits of the same dataset, we conclude that functional heterogeneity may be a dominating factor behind the lack of reproducibility of FC alterations in different rs-fMRI studies of PD. While global PD-related functional connectivity changes were non-reproducible across datasets, we identified a few individual brain region pairs with marginally consistent FC changes across all three datasets. However, training classifiers on each one of the 3 datasets to discriminate PD scans from controls produced only low accuracies on the remaining two test datasets. Moreover, classifiers trained and tested on random splits of the same dataset (which are technically homogeneous) also had low test accuracies, directly substantiating disease heterogeneity.
q-bio.NC
q-bio
Exploring the reproducibility of functional connectivity alterations in Parkinson's Disease Liviu Badea1*, Mihaela Onu2,3, Tao Wu4,5, Adina Roceanu6, Ovidiu Bajenaru6,7 1 Artificial Intelligence and Bioinformatics Group, National Institute for Research and Development in Informatics, Bucharest, Romania 2 Medical Imaging Department, Clinical Hospital Prof. Dr. Th. Burghele, 20 Panduri Street, Bucharest, Romania 3 University of Medicine and Pharmacy "Carol Davila", Biophysics Department, Bucharest, Romania 4 Department of Neurobiology, Key Laboratory on Neurodegenerative Disorders of Ministry of Education, Beijing Institute of Geriatrics, Xuanwu Hospital, Capital Medical University, Beijing, China 5 Beijing Key Laboratory on Parkinson's Disease, Parkinson Disease Centre of Beijing Institute for Brain Disorders, Beijing, China 6 University Emergency Hospital Bucharest, Neurology Department, Bucharest, Romania 7 University of Medicine and Pharmacy "Carol Davila", Department of Clinical Neurosciences, Bucharest, Romania * Corresponding author Abstract Since anatomic MRI is presently not able to directly discern neuronal loss in Parkinson's Disease (PD), studying the associated functional connectivity (FC) changes seems a promising approach toward developing non-invasive and non-radioactive neuroimaging markers for this disease. While several groups have reported such FC changes in PD, there are also significant discrepancies between studies. Investigating the reproducibility of PD-related FC changes on independent datasets is therefore of crucial importance. We acquired resting-state fMRI scans for 43 subjects (27 patients and 16 normal controls, with 2 replicate scans per subject) and compared the observed FC changes with those obtained in two independent datasets, one made available by the PPMI consortium (91 patients, 18 controls) and a second one by the group of Tao Wu (20 patients, 20 controls). Unfortunately, PD-related functional connectivity changes turned out to be non-reproducible across datasets. This could be due to disease heterogeneity, but also to technical differences. To distinguish between the two, we devised a method to directly check for disease heterogeneity using random splits of a single dataset. Since we still observe non-reproducibility in a large fraction of random splits of the same dataset, we conclude that functional heterogeneity may be a dominating factor behind the lack of reproducibility of FC alterations in different rs-fMRI studies of PD. While global PD-related functional connectivity changes were non-reproducible across datasets, we identified a few individual brain region pairs with marginally consistent FC changes across all three datasets. However, training classifiers on each one of the three datasets to discriminate PD scans from controls produced only low accuracies on the remaining two test datasets. Moreover, 1 classifiers trained and tested on random splits of the same dataset (which are technically homogeneous) also had low test accuracies, directly substantiating disease heterogeneity. Keywords: Reproducibility; Resting state functional connectivity markers; Parkinson's Disease; Data sharing Abbreviations PD Parkinson's Disease NC Normal Controls rs-fMRI resting state functional Magnetic Resonance Imaging FC Functional Connectivity ROI Region of Interest H&Y Hoehn and Yahr score PPMI Parkinson's Progression Markers Initiative SVM Support Vector Machine classifier GNB Gaussian Naïve Bayes classifier 2 Introduction Although Parkinson's disease (PD) is the second most common neurodegenerative disease after Alzheimer's disease, its diagnosis is still difficult, especially in the early premotor stages, as it is mainly based on clinical evidence. To date, there is still no unique standard diagnostic test for PD, despite the intense research efforts to develop accurate biomarkers based on blood tests or imaging scans. The best current objective tests for PD evaluate dopaminergic function in the basal ganglia by using various PET or SPECT radiotracers (e.g. DaTSCAN). But these tests make use of radioactive substances, are performed only in specialized imaging centers and can also be very expensive. Moreover, the loss of dopaminergic nigro-striatal neurons is a delayed pathological event in the evolution of the disease, corresponding to Braak stages III-IV. On the other hand, conventional (CT or MRI) brain scans of PD patients usually appear normal or with minor non-specific changes, so that conventional imaging techniques are only useful for ruling out other diseases that can be secondary causes of parkinsonism. Therefore, since anatomic MRI is presently not able to directly discern (dopaminergic) neuronal loss in PD [Tuite et al., 2013], studying the associated functional connectivity (FC) changes seems to be a promising approach toward developing non-invasive and non-radioactive neuroimaging markers for this disease. While many groups have reported such FC changes in PD (see Supplementary Table 1 for a list of such studies), an in-depth analysis of existing literature revealed significant discrepancies between studies. Investigating the reproducibility of PD FC changes on independent datasets is therefore of crucial importance. A comprehensive review and analysis of the literature related to resting-state fMRI studies of Parkinson's disease is out of the scope of the present paper [Gottlich et al., 2013; Long et al., 2012; Skidmore et al., 2013; Baudrexel et al., 2011; Wu et al., 2011; Kwak et al., 2010; Kwak et al., 2012; Helmich et al., 2010; Helmich et al., 2011; Luo et al., 2014; Yu et al., 2013; Hacker et al., 2012; Kurani et al., 2015; Baggio et al., 2015; Esposito et al., 2013; Szewczyk-Krolikowski et al., 2014; Tessitore et al., 2012; Sharman et al., 2013; Wu et al., 2012; Liu et al., 2013; Chen et al., 2015; Wen et al., 2013; Zhang et al., 2015] (Supplementary Table 1; see also the review by [Tahmasian et al., 2015]). We only mention some important inconsistencies of reported functional connectivity changes in PD. Due to the crucial importance of the striatum in PD, we first discuss some inconsistencies involving striatal seeds [Tahmasian et al., 2015]:  Contrary to [Hacker et al. 2012], [Helmich et al. 2010] observed no significant difference in caudate functional connectivity in PD.  On the other hand, contrary to the study [Helmich et al., 2010], [Luo, Song, et al. 2014] did not observe increased FC of the anterior putamen.  In contrast to [Hacker et al., 2012], [Luo, Song, et al. 2014] did not find a FC decrease between the striatal seeds and the brainstem. There are also inconsistencies involving non-striatal seeds. For example, [Wu et al., 2011] found disrupted FC between the pre-SMA and the left putamen, as opposed to [Helmich et al., 2010], who did not find a decreased FC between the putamen and pre-SMA in PD. 3 Since the motor symptoms are the most striking clinical manifestations in PD, many rs-fMRI studies of PD concentrate on the sensorimotor system, including the basal ganglia, while disregarding any other FC changes. On the other hand, other more unbiased studies tried to determine a more global picture of the FC changes in PD. Some even tried to develop classifiers for the disease based on rs-fMRI data [Long et al., 2012; Skidmore et al., 2013; Szewczyk- Krolikowski et al., 2014; Chen et al., 2015], but most studies were not validated on independent datasets. There are also some gross discrepancies involving even the sign of the main FC changes in PD. For example, [Luo et al. 2014] found only decreased FC in early stage PD, whereas most studies also find FC increases. The general picture one gets from the literature is complex and at times somewhat confusing due to the numerous inconsistencies. Of course, these inconsistencies could be due to the different disease stages analyzed, to the inherent functional heterogeneity of the disease, but also to technical differences, or to the differences in the complex data (pre)processing workflows. Therefore, it is crucial to use the same data processing workflow to check the reproducibility of PD-related FC changes on as many independent datasets as possible. In this work, we report a comparison between three different datasets obtained by completely independent research groups (see Table 1). More precisely, we acquired resting-state scans for 43 Romanian subjects (27 patients and 16 normal controls, with 2 replicate scans per subject) and compared the observed functional connectivity changes with those obtained in two independent datasets, one made available by the PPMI consortium in the US (91 patients, 18 controls) and a second one by the group of Tao Wu in China (20 patients and 20 normal controls). Table 1. The 3 PD datasets compared in the present study (PD=Parkinson's Disease, NC=normal controls). PD subjects NC Dataset subjects NEUROCON 27 (16 M) 16 (5 M) Tao Wu PPMI PD scans 54 20 (11 M) 20 (12 M) 20 91 (63 M) 18 (14 M) 134 NC scans 31 20 19 age (NC) mean (SD) age PD mean(SD) 68.7 (10.6) 67.6 (11.9) 0.76 1.92 (0.33) 4.6 (6.5) 65.2 (4.4) 64.8 (5.6) 0.78 1.88 (0.63) 5.4 (3.9) 61.3 (10.2) 64.7 (9.7) 0.17 1.72 (0.48) 1.9 (1.0) p (age NC- PD) Hoehn & Yahr mean(SD) disease duration mean(SD) We briefly describe our approach in Fig 1A to better guide the reader through the remainder of the paper. We started by comparing the PD-related global functional connectivity changes in the 3 datasets and found them to be non-reproducible. Of course, this could be due to disease heterogeneity, but also to technical differences. To better distinguish between these two possibilities, we devised a method to directly check for disease heterogeneity using random splits of a single dataset. On the other hand, we searched for individual brain region pairs with consistent connectivity changes across all three datasets. Finally, to more directly discriminate PD scans from controls, we trained multivariate machine learning classifiers on one dataset and tested them on the remaining two. Additionally, we trained and tested classifiers on technically homogeneous random splits of the same dataset, to more directly check for disease heterogeneity. 4 Fig 1. Overview. (A) Main steps of the analysis. (B) Using random splits of a dataset with replicate scans to check for disease (group) heterogeneity: (a, right) by placing different subjects (with all their replicate scans) in the two splits ("split subjects") and respectively (b, left) by splitting the replicates of the same subjects in the two splits ("split replicates"). The "split replicates" datasets must show reproducible changes anyway, while non-reproducible changes across the "split subjects" datasets are an indication of disease heterogeneity. (a,b,c,... correspond to subjects, while, for instance, a' and a" are replicate scans for subject a.) 5 This is the first study investigating the reproducibility of functional connectivity changes in Parkinson's disease on more than 2 datasets. Given the paucity of publicly available rs-fMRI PD datasets, we advocate the critical importance of data sharing for enabling the discovery of reproducible rs-fMRI biomarkers of PD. Materials and Methods Datasets Three resting-state fMRI datasets of Parkinson's disease were compared in this study (see also Table 1 for the main patient characteristics and the numbers of scans in each study): - - - the NEUROCON rs-fMRI study of 27 PD patients and 16 normal controls of the Neurology Department of the University Emergency Hospital Bucharest (Romania), a dataset of 20 PD patients and 20 normal controls provided by the group of Tao Wu (China), a publicly available dataset of 91 PD patients and 18 controls of the Parkinson's Progression Markers Initiative (PPMI) study in the US. The datasets are somewhat similar, except for PPMI, which involved patients with a diagnosis of PD for two years or less and who are not taking PD medications, while most patients from the other two studies have been under treatment (most under levodopa). Also, PPMI patients were scanned in the 'eyes open' condition. Still, we argue that our findings were not affected by these differences. Since the datasets were compared in a pairwise manner, any putative discrepancies due to the shorter disease durations in the PPMI dataset would only show up in the NEUROCON-PPMI and Tao Wu-PPMI comparisons, but not in the NEUROCON-Tao Wu comparison. This was not observed in reality. NEUROCON The NEUROCON study enrolled 27 patients with Parkinson's disease (mean age±SD 68.7±10.6 years) and 16 age-matched normal controls (67.6±11.9 years) with no history of neurological or psychiatric disease. The patients were clinically assessed at the Neurology Department of the University Emergency Hospital Bucharest (Romania) to be in the early or moderate stage of the disease according to the Queen Square Brain Bank (QSBB) clinical criteria and met the EFNS/MDS-ES (European Federation of Neurological Societies/Movement Disorder Society– European Section) recommendations for diagnosis of Parkinson's disease. The mean disease duration was 4.6 (±6.5) years for the entire patient cohort and respectively 2.75 (±2.15) years after excluding three patients with particularly long disease durations (over 10 years: 11, 16 and 32 years, respectively). Despite the longer disease durations, the above-mentioned 3 patients met the criteria for moderately advanced disease (H&Y stage 2) and thus were included in the study. The mean Hoehn and Yahr (H&Y) score [Hoehn and Yahr, 2001] was 1.93 (±0.33) and respectively 1.92 (±0.35) after excluding the 3 patients with long disease durations. All patients were in an early to moderate stage of disease (stages 1 to 2.5). The mean score on the motor subset of the Unified Parkinson's Disease Rating Scale (UPDRS) [Fahn, 1986] in the off medication condition was 28.3 (±9.3) for the entire patient cohort and 26.9 (±8.8) after excluding the 3 patients with long disease 6 durations. The study has been approved by the University Emergency Hospital Bucharest ethics committee in accordance with the ethical standards of the 1964 Declaration of Helsinki and its later amendments. All patients gave their written informed consent to participate in the study. Scanning. All subjects underwent two consecutive 8 min fMRI scans in a 1.5-Tesla Siemens Avanto MRI scanner, in an awake resting state with their eyes closed. Two consecutive replicate scans were acquired for each subject to enable the study of the reproducibility and respectively homogeneity of FC changes in PD. (A single (control) subject could only be scanned once.) The patients were scanned in the "off medication" state, at least 10 hours after the last intake of their medication. The scanning protocol involved an Echo Planar sequence with repetition time (TR) 3480 ms, echo time (TE) 50 ms, axial orientation, voxel size 3.8×3.8×5 mm (without slice gaps), flip angle 90 and number of averages=1. Each resting state session lasted 8.05 min, comprising 137 volumes. To enable better co-registration to the standard MNI template, high-resolution T1- weighted images were also obtained for all subjects using an MPRAGE sequence (IR method, TR=1940ms, TE=3.08ms, inversion time (IT)=1100ms, voxel size 0.97×0.97×1 mm, number of averages=1). Tao Wu The dataset comprised 20 PD patients (11 males, mean age±SD 65.2±4.4 years) and 20 age- matched normal controls (12 males, 64.8±5.6 years). All patients were in the early to moderate stage of the disease (Hoehn and Yahr stages 1 to 2.5, except for a single patient with H&Y stage 3) and had normal Mini-Mental State Examination (MMSE) scores. Moreover, there was no statistically significant MMSE difference (p=0.43) between PD patients (28.8±1.1) and normal controls (29.1±1.3). Scanning. Both resting state fMRI and anatomic T1 scans were acquired for the 40 subjects in a Siemens Magnetom Trio 3T equipment, in an awake resting state with their eyes closed. Each resting state session lasted about 8 min (239 volumes, TR=2s, TE=40 ms, flip angle=90) with a voxel size of 4×4×5 mm (6464 matrix, 28 slices, field of view=256mm×256mm). MPRAGE scans were also obtained (voxel size 1×1×1 mm) for registration to the MNI template. PPMI Imaging data for 91 PD patients (63 males) and 18 normal controls (14 males) were downloaded from the Parkinson's Progression Markers Initiative (PPMI) study data portal (http://www.ppmi- info.org/access-data-specimens/download-data/, https://ida.loni.usc.edu/home/projectPage.jsp?project=PPMI) with the kind permission of the PPMI consortium. The study includes patients with a diagnosis of PD for two years or less and who are not taking PD medications. The patients and controls are age-matched (p=0.17, mean age±SD 61.3±10.2 years for the PD patients and respectively 64.7±9.7 years for the normal controls). The patients had a mean Hoehn & Yahr score of 1.72 (SD 0.48), with a mean disease duration (at the time of the scan) of 1.9 years (SD 1.0). All patients had H&Y scores 1 to 2, except for only two, who were classified as H&Y stage 3. Scanning. The subjects were scanned in 8 different centers, but with a similar protocol on Siemens Tim Trio 3Tesla scanners. Each resting state session lasted about 8.4 min (210 volumes, TR=2.4s, TE=25ms, flip angle 80) with a voxel size of 3.3×3.3×3.3 mm (6866 matrix, 40 slices). Subjects were instructed to rest quietly, keeping their eyes open and not to fall asleep. MPRAGE scans were 7 also obtained (voxel size 1×1×1 mm, TR=2.3s, TE=2.98ms, flip angle=9) for registration to the MNI template. Although our functional connectivity computations did not require any particular type of data normalization (as only inter-region correlations are computed, rather than amplitudes), we also considered a subset of scans acquired in a single center (center number 32, with the largest number of PD patient and normal control scans), referred to in the following by the suffix 'center32'. Preprocessing to NIfTI using dcm2nii (https://www.nitrc.org/projects/dcm2nii/) and All datasets were preprocessed in a uniform manner. The raw scanner data in DICOM format was converted further preprocessed using FSL (FMRIB Software Library v5 http://fsl.fmrib.ox.ac.uk/fsl/) as follows: motion correction using MCFLIRT, brain extraction with BET, spatial smoothing (Gaussian kernel FWHM 5mm) and denoising using nonlinear filtering (SUSAN), temporal high-pass filtering (with a cutoff frequency of 0.01 Hz), registration to the standard Montreal Neurological Institute MNI152 template via the anatomical T1 image (more precisely, BBR registration of the BOLD image to the T1 image, followed by 12 DOF linear+nonlinear registration of the latter to the 2mm MNI template). Nonlinear registration was performed at a resampling resolution of 4mm. Besides the above 'standard' preprocessing workflow, we also considered alternative workflows involving global signal regression (GS) and respectively a temporal bandpass filter (0.01-0.1Hz – an ideal low-pass 0.1Hz filter was used in addition to the default FSL 0.01Hz highpass filter). Since subject motion in the scanner has been observed to have significant influence on the functional connectivities computed from rs-fMRI data, despite motion correction (e.g. [Power et al., 2015]), we also considered subsets of scans with low in-scanner motion (marked by the suffix '0', e.g. 'NC0' and 'PD0' – see also Supplementary Table 2). PD-related functional connectivity changes Functional connectivity [Friston and Buchel, 2003] is a rather loosely defined term, which encompasses many different methods used to reveal temporal correlations of BOLD activity across the brain. The simplest method consists in computing the correlations between all pairs of regions of a given brain parcellation, but more sophisticated data decomposition methods, such as Independent Component Analysis (ICA) are also widely used. Such data decomposition methods do not assume a given brain parcellation, but instead construct spatial maps grouping voxels with highly correlated timecourses. (Still, instead of being given a parcellation, such methods need to be provided with a target number of components.) In our study of the reproducibility of functional connectivity changes in PD, we used brain parcellations constructed independently of the datasets under comparison, rather than applying data decomposition methods, such as group-ICA, since the latter would be inherently biased towards the "training dataset". Group-ICA may obtain a better functional parcellation for the training dataset, but that parcellation would be less appropriate for any other independent dataset ("overfitting"), thereby introducing a bias in the analysis. To avoid these problems, we have chosen to use brain parcellations constructed independently of the datasets under comparison, including functional brain parcellations obtained by group-ICA on completely independent sets of subjects (such as the 'Stanford' functional parcellation [Shirer et al., 2011]). 8 Moreover, to compensate for potential biases of any specific parcellation, we extended our analyses to a number of 13 different parcellations employed in other rs-fMRI studies, two anatomical (AAL, Talairach) and 11 functional (see Table 2 for more details). Table 2. The brain parcellations used in the functional connectivity comparisons. Number of regions Comments Parcellation Reference AAL [Tzourio-Mazoyer et al., 2002] [Craddock et al., 2012] Craddock 130 Craddock 260 Craddock 500 Craddock 950 Shen 100 Shen 200 Shen 300 OASIS Power Gordon_surface [Gordon et al., 2014] [Marcus et al., 2007] [Power et al., 2011] [Shen et al., 2013] Talairach [Talairach and Tournoux, 1988] Stanford [Shirer et al., 2011] 90 116 130 260 500 950 93 183 278 97 264 333 695 anatomic atlas spherical regions with a 10mm radius anatomic atlas functional parcellation obtained by group-ICA For each parcellation, we computed average timecourses for each region of interest (ROI) and the resting state functional connectivity between each pair of ROIs (ROI1,ROI2) as the Fisher z- transform of the temporal correlation between the corresponding ROI timecourses: FC(ROI1,ROI2) = z(corr(ROI1,ROI2)). For each dataset, we determined significant PD-related FC changes by applying two-sample t-tests (with unequal sample sizes and unequal variances) to the functional connectivities of all ROI pairs. ROI-pairs with significant group differences (NC versus PD) represent regions whose functional connectivity was found to be significantly different in PD patients in that particular dataset. The main aim of this study is to determine whether these changes are reproducible across datasets, to enable the development of functional imaging biomarkers for PD. Reproducibility of global functional connectivity changes in PD Comparison of 3 different PD datasets We first compared the global PD-related functional connectivity changes across the three independent datasets NEUROCON, Tao Wu and PPMI to check to what extent these changes are reproducible. More precisely, we performed pairwise comparisons for all dataset pairs as follows. For each pair of datasets (i,j) and a fixed parcellation, we checked the extent to which the PD- related FC changes in one dataset are correlated to the changes in the second dataset. 9 PD-related FC changes were quantified using t-values t(ROIk,ROIl) from group comparisons (unpaired two-sample t-tests between NC and PD) of the functional connectivities between pairs of regions of interest FC(ROIk,ROIl). Then the reproducibility Rij across the two datasets i and j was determined as the correlation between the corresponding t-values (viewed as a vector over all ROI pairs) for the two datasets: where Rij = corr(Ti,Tj), (1) Ti=(ti(ROI1,ROI2), ti(ROI1,ROI3), ti(ROI1,ROI4), ) with ti(ROIk,ROIl) the t-value corresponding to PD-related FC changes between ROIk and ROIl with respect to dataset i (and similarly Tj for dataset j). For a more intuitive graphical depiction of reproducibility across two datasets, we also constructed the scatter-plot of ROI-pair t-values corresponding to group comparisons in the two datasets (see Fig 2 for an example of such a scatter-plot). Comparing PD-related FC changes (t-values) in the two datasets amounts to plotting for each ROI- pair the t-value in dataset 1 against the t-value in dataset 2. We thereby obtain a scatter-plot with a point for each ROI pair. The comparison of the FC changes in the two datasets thus involves analyzing the distribution of points in the scatter-plot: ideally, perfect reproducibility would entail a diagonal distribution of points in the scatter-plot, corresponding to perfectly correlated t-values in the two datasets. Fig 4 depicts examples of good reproducibility, while Fig 2 shows cases of non- reproducibility across datasets. The correlation of t-values for the two datasets Rij = corr(Ti,Tj), as introduced above in (1), can be viewed as an aggregate measure of the reproducibility across the two datasets i and j. To obtain a more quantitative measure of the statistical significance of such a correlation Rij between datasets, we performed permutations of the group labels (NC and PD) independently for the two datasets and computed the p-value of the Rij value as the fraction of permutations  for which the dataset correlation w.r.t. the permuted data Rij pij = { permutation  Rij () exceeds the real one (Rij): ()  Rij } / N, (2) where N is the total number of permutations. All our permutation tests involved N=1000 permutations. Various factors have been mentioned in the literature to affect functional connectivity measures: - subject motion in the scanner [Power et al., 2015], - global signal regression (with or without) [Murphy et al., 2009; Hayasaka 2013], - the choice of the parcellation. To study the influence of these factors on our reproducibility results, we also considered subsets of scans with low in-scanner motion (marked by the suffix '0', e.g. 'NC0' and 'PD0'), repeated our analyses with global signal regression, bandpass filtering and performed the comparisons using all 13 brain parcellations previously mentioned. Since the PPMI data has been acquired in several different imaging centers, we also considered a potentially more homogeneous subset of scans 10 acquired in a single center (center number 32, with the largest number of PD patient and normal control scans), referred in the following by the suffix 'center32'. Comparison of random splits of the same PD dataset As already mentioned in the Introduction, the observed lack of reproducibility of global FC changes across datasets could be due to disease heterogeneity, but also to technical differences. To distinguish between these two possibilities, we devised a method to directly check for disease heterogeneity using random splits of a single dataset with replicate scans. Technical differences can then be excluded since all the scans have been acquired under identical technical conditions. More precisely, since in the NEUROCON study we acquired two replicate scans for each subject, we constructed two homogeneous dataset splits simply by using (different) scans of the same subjects. Additionally, two heterogeneous dataset splits can be obtained by placing different subjects (with all their scans) in each split. In other words, instead of comparing two distinct datasets, we compared two random splits of the same dataset, either: (a) by placing different subjects in the two splits, with all the replicate scans of a subject in the same split ("split subjects", heterogeneous split), or (b) by placing each replicate scan of the same subject in a different split, so that the two splits contain (different) scans of the same subjects ("split replicates", homogeneous split). Dataset splits (b) are homogeneous since they contain scans of the same subjects, while splits (a) they contain scans of different subjects. Therefore, consistent are heterogeneous since reproducibility across all random heterogeneous splits would indicate disease homogeneity, while non-reproducibility in a large fraction of random heterogeneous splits would imply disease heterogeneity. (In both cases, we expect to observe consistent reproducibility across the homogeneous splits, at least as long as the technical noise is not dominating the biological signal.) A diagram of our method is shown in Fig 1B. As in the pairwise comparisons between different datasets, we used permutation tests and formula (2) to compute p-values of the reproducibility across split datasets, for both the heterogeneous ("split subjects") and the homogeneous ("split replicates") datasets. Due to the random nature of the splits, we repeated the analysis for Ns > 1000 different random splits of the original data. To assess the fraction of (non-)reproducible splits, we determined the empirical cumulative distribution function (CDF) of the reproducibility p-values for the Ns random splits. The analysis was also repeated for the data with global signal regression. Influence of technical factors, preprocessing and parcellation We also studied the influence on reproducibility of certain key technical factors and preprocessing steps, such as: - - the repetition time (TR), linear vs. nonlinear registration, - global signal regression, - the specific brain parcellation used for evaluating functional connectivity. The AAL parcellation (which is typical) was used whenever not specified otherwise. 11 Doubling the TR The repetition time might, in principle, influence the measured low-frequency rs-fMRI fluctuations and indirectly the functional connectivities (which are temporal correlations). To study the influence of the repetition time on reproducibility, we constructed a synthetic dataset with a double TR by leaving out every second time-point from the NEUROCON timeseries data (for each scan and each voxel). We then analyzed with our method the reproducibility of group changes in functional connectivity between the original NEUROCON dataset and the synthetic one with a double TR. Linear vs. nonlinear registration To study the impact of registration on reproducibility, we preprocessed the NEUROCON data both with linear and nonlinear registration to the MNI 152 template and determined the reproducibility of group changes in functional connectivity between the two resulting datasets. Global signal regression Since global signal regression (GSR) has been observed to be very effective at removing scanning artifacts [Hayasaka, 2013], including motion artifacts, we also studied the reproducibility of FC changes between the NEUROCON dataset processed with GSR and the same dataset processed without GSR. Influence of parcellation To avoid potential biases of any specific parcellation, we repeated our pairwise comparisons between the 3 PD datasets using all 13 different parcellations mentioned above, including functional and anatomic parcellations, with a wide range of numbers of regions of interest (90 to 950). Reproducibility of the individual differentiating FC changes The reproducibility analysis performed above involves global functional connectivity changes, i.e changes in the FC of all ROI pairs, not just the ones that differentiate PD from normal controls. Even with non-reproducible global FC changes, it might be in principle possible that only a very few brain region pairs might still reproducibly differentiate PD patients from controls. To this end, we also studied individual brain region pairs with FC changes that are significant w.r.t. all datasets. More precisely, for each ROI-pair, we compute max(p), the largest (least significant) of the three p- values obtained in the three datasets (separately for the FC increases and respectively decreases) and sort the ROI-pairs in increasing order of this max(p). The most significant min(max(p)) of these max(p) corresponds to the ROI-pair with the best overall significance with respect to the 3 datasets, as all other ROI-pairs have larger (less significant) p-values with respect to at least one dataset. Finally, to assess the statistical significance of such a best ROI-pair, we use a permutation test (of the disease labels in each dataset) to check the fraction of random permutations with a more significant (smaller) min(max(p)) than the real data (we performed N=1000 random permutations). p(min(max(p+))) = { permutation  min(max(p+))()  min(max(p+)) } / N, where min(max(p+)) corresponds to FC increases in NC versus PD. A similar relation holds for the FC decreases min(max(p)). The analysis was repeated for all 13 parcellations considered in this study (Table 2). 12 Learning classifiers for discriminating PD-related FC changes We used machine learning techniques to learn classifiers that discriminate PD from controls using functional connectivities between ROI pairs as features. First, we trained classifiers on each one of the 3 datasets (NEUROCON, Tao Wu, PPMI) and tested them on the other two datasets. Both Linear Support Vector Machines (SVM) and Gaussian Naïve Bayes (GNB) classifiers were tested, with progressively increasing numbers of features: N=10,50,100,500,5000 (functional connectivities between ROI pairs). The N features selected were the best discriminating ROI pair functional connectivities, based on unpaired t-tests between normal and PD scans. As the two classes (NC-Normal Controls and PD-Parkinson's Disease) are not balanced in all 3 datasets, we employ the average accuracy Aacc = (acc(NC)+acc(PD))/2 for assessing the performance of the classifiers (a random classifier is expected to have an average accuracy of 0.5). Since the different datasets are not technically homogeneous, we also trained and tested classifiers on random splits of the same dataset, to check to what extent the low accuracies are due to technical differences, or to disease heterogeneity. More precisely, we performed 10,000 random splits in half of each dataset, trained a classifier on one half and tested it on the other. 13 Results ROI-pairs with significant group differences (NC versus PD) in functional connectivity were found in all three PD datasets: NEUROCON, Tao Wu and PPMI (Supplementary Tables 3 and 4). However, these changes seemed at first sight to be distinct in each dataset. Our main aim in this paper has been to systematically investigate the reproducibility of the PD-related FC changes across independent validation datasets. PD-related FC changes are non-reproducible across 3 datasets The reproducibility of global PD-related functional connectivity changes was determined by pairwise comparisons between three independent datasets: NEUROCON, Tao Wu and respectively PPMI. Fig 2 shows the scatter-plots of ROI-pair t-values (corresponding to the group comparison NC-PD) for the three dataset pairs, indicating a lack of reproducibility of global FC changes in PD. (Perfect reproducibility would correspond to a diagonal distribution of points corresponding to ROI pairs with perfectly correlated t-values with respect to both datasets.) Moreover, discriminating ROI-pairs situated in the upper right and respectively lower left corners of one plot are not discriminating in the other plots. Fig 2. Scatter-plots of ROI-pair t-values for the three dataset pairs indicate non- reproducibility of global PD-related FC changes. For a more quantitative measure of the reproducibility of FC changes between two datasets, we computed the Pearson correlation between t-values (viewed as vectors of over all ROI pairs) with respect to each dataset (R values shown in Fig 2.) We also estimated the statistical significance (p- values) of these reproducibility measures by permutation tests of the group labels independently for the two datasets – Table 3 shows the reproducibility measure and associated p-value for various pairwise comparisons between the three datasets, with standard highpass preprocessing (>0.01Hz), global signal regression (GS) and respectively bandpass filter (0.01-0.1Hz). Since in-scanner motion may influence FC measures, we present not only a comparison between the full patient and normal control cohort, but also that corresponding to a subset of scans with low in-scanner motion 14 (denoted by the suffix '0'). Moreover, since PPMI data were acquired at many different centers, we also considered the restriction of the PPMI data to the scans from a single center (suffix 'center32'). The AAL parcellation was used in this case, but we also study the influence of the parcellation later on. Table 3. Reproducibility measure R and associated p-value p for various pairwise comparisons between datasets with standard highpass preprocessing ('standard'), global signal regression (GS) and respectively bandpass filter (BP) ('0' indicates 'low in scanner motion', 'center32' – scans performed at a single PPMI center). Dataset 1 NEUROCON Group contrast 1 NC-PD NC0-PD0 NC-PD NC0-PD0 Dataset 2 Tao Wu NEUROCON NC-PD PPMI NC0-PD0 NC-PD NC0-PD0 Tao Wu NC-PD PPMI NC0-PD0 Group contrast 2 NC-PD NC0-PD0 NC-PD NC0-PD0 NC_center32- PD_center32 NC0_center32- PD0_center32 NC-PD NC0-PD0 NC_center32- PD_center32 NC0_center32- PD0_center32 standard R p GS BP R p R p -0.145955 0.949 -0.0679698 0.804 0.000845579 0.476 -0.0832455 0.847 0.035725 0.300 0.0132475 0.437 0.0692377 0.257 -0.163251 0.937 0.0401831 0.352 0.0308243 0.373 -0.0824776 0.800 0.10039 0.138 -0.0122797 0.526 0.0188561 0.430 -0.142598 0.914 -0.0873994 0.813 -0.0894912 0.832 -0.152545 0.945 0.00673472 0.466 0.0228885 0.403 0.0948706 0.153 -0.0182896 0.560 0.044843 0.298 0.0545817 0.293 -0.0446155 0.687 -0.032841 0.666 0.0243423 0.400 -0.0712155 0.812 -0.0176842 0.597 -0.0436096 0.681 A clear lack of reproducibility of global PD-related FC changes is observed in all the three dataset pairs. This is the first study comparing three independent rs-fMRI datasets of PD. The fact that we compare 3 datasets is very important, as it lowers the probability that the lack of reproducibility is due to a dataset that may be "faulty" in some sense – in that case, with 3 datasets we might still observe reproducibility with respect to the remaining dataset pair (which we do not see in reality). Inconsistent reproducibility of FC changes in heterogeneous dataset splits indicate disease heterogeneity The non-reproducibility across 3 datasets mentioned above seems to be due to disease heterogeneity, but it could also be due to technical differences. To exclude the latter possibility, we checked for disease heterogeneity using random splits of a single dataset with replicate scans (NEUROCON), all of which have been acquired under identical technical conditions. (a) We first constructed random splits by placing different subjects in the two splits, with all the replicate scans of a subject in the same split ("split subjects", heterogeneous splits). As can be seen in Fig 3A (blue curve), a large fraction of these random splits display non-reproducible functional connectivity changes. More precisely, 88% of the random heterogeneous splits show non- reproducibility at the p>0.01 level and 42% at the p>0.05 level. Fig 3B shows the complementary cumulative distribution function (1-CDF) for the corresponding reproducibility measure R (blue curve), while Fig 3C presents a typical scatter-plot of ROI-pair t-values for a random heterogeneous split (one with R equal to the median). 15 Fig 3. Inconsistent reproducibility of PD-related FC changes in random heterogeneous dataset splits and consistent reproducibility in random homogeneous dataset splits. (A) Complementary cumulative distribution function (CCDF=1-CDF) of the reproducibility p-values for Ns=2510 random heterogeneous splits and Ns=325 random homogeneous splits. (B) CCDF of the reproducibility measure R. (C) A scatter-plot of ROI-pair t-values for a random heterogeneous split. (D) A scatter-plot of ROI-pair t-values for a random homogeneous split. (b) Next, we constructed random splits by placing each replicate scan of the same subject in a different split, so that the two splits contain (different) scans of the same subjects ("split replicates", homogeneous splits). In contrast to (a), all homogeneous splits showed reproducibility at the p<10-3 level (Fig 3A, red curve). Fig 3B shows the complementary CDF for the reproducibility measure (red curve) – note the significantly higher reproducibility (R) values for the homogeneous splits (red curve, median R=0.71) as compared to the heterogeneous splits (blue curve, median R=0.21). Fig 3D displays a typical scatter-plot of ROI-pair t-values for a random homogeneous split. The observed non-reproducibility in a large fraction of heterogeneous dataset splits indicates disease heterogeneity, in line with the comparison between the 3 independent PD datasets. As a control, we did indeed observe consistent reproducibility with respect to all homogeneous dataset splits, demonstrating that the technical noise could not have been the dominating factor behind the erratic non-reproducibility in heterogeneous splits. 16 The fact that the well-known clinical heterogeneity of Parkinson's disease is also accompanied by heterogeneity in resting state functional connectivity may not retrospectively be a big surprise to an experienced neurologist, although its exact extent could not have been estimated a priori, before analyzing the data. However, does this FC heterogeneity in PD also imply the lack of practical usefulness of rs-fMRI functional connectivity? Are there any other conditions that can be reliably differentiated using resting state functional connectivity? To answer these questions, we applied our approach to a different, potentially more homogeneous contrast, namely that between eyes open and eyes closed resting state conditions in healthy volunteers. Repeating our analysis of reproducibility of FC group changes on random splits of the Beijing eyes open-eyes closed dataset [Liu et al., 2013] (see Supporting Information) revealed reproducibility (p<0.05) not just in the homogeneous dataset splits, but also in the heterogeneous ones (Supplementary Fig 1 – only 6% of the heterogeneous and just 0.8% of the homogeneous random splits were non-reproducible at the p>0.05 level). Summing up our findings, from the point of view of global FC changes, Parkinson's disease is heterogeneous, as opposed to the eyes open-eyes closed contrast, which is much more homogeneous (Table 4). Table 4. Summary of reproducibility of global functional connectivity changes in Parkinson's Disease and respectively the Eyes Open-Eyes Closed contrast. Reproducibility of global FC changes different datasets PD no split subjects (heterogeneous splits) inconsistent (no in a large fraction of splits) split replicates (homogeneous splits) yes EO-EC yes yes Influence of technical factors and preprocessing on reproducibility We found good reproducibility when changing various technical factors or processing options of the NEUROCON data, such as (see Fig 4): - doubling the repetition time (TR), - registration (linear versus nonlinear), - global signal regression (with versus without). 17 Fig 4. Reproducibility when changing various technical factors or preprocessing options. This is in line with our conclusion that functional heterogeneity, rather than these technical factors, is the dominating factor behind the lack of reproducibility of FC changes in different rs-fMRI studies of Parkinson's disease. We also tested the influence of various rs-fMRI denoising methods on the reproducibility of PD- related FC changes, such as ICA-FIX [Griffanti et al., 2014; Salimi-Khorshidi et al., 2014], or regression of the mean white matter and/or cerebrospinal fluid signal – none of these denoising methods changed the observed non-reproducibility (data not shown). We also observed no improvement in reproducibility across random splits of the NEUROCON dataset after regressing out potential confounders, such as age, gender, or disease duration (data not shown). Influence of parcellation on reproducibility We have argued that functional connectivity must be computed with respect to an unbiased parcellation (i.e. one that hasn't been constructed from any of the analyzed datasets). However, any given parcellation has also specific biases that may in principle affect the capacity to discriminate between PD and normal controls – an especially relevant factor is the average size and number of the ROIs. Testing the reproducibility of the PD-related global FC changes using 13 different parcellations, with varying numbers of ROIs (between 90 and 950, see Table 2), revealed a lack of reproducibility regardless of parcellation, or dataset pair (Table 5). (The NEUROCON-PPMI comparison was marginally significant (p=0.05) for the NC-PD contrast, but this significance didn't survive perturbations such as selecting just the 'center32' scans from PPMI (p=0.259), or restriction to the low motion scans NC0-PD0 (p=0.26), or NC0-PD0_center32 (p=0.555).) 18 Table 5. Reproducibility measure and associated p-value for 13 parcellations and all three dataset pairs (NC-PD contrast). NEUROCON-TaoWu NEUROCON-PPMI TaoWu-PPMI R p R p R -0.011116 -0.0490741 -0.032232 -0.0247647 0.172 -0.0546894 0.554 0.0927652 0.191 -0.0613781 -0.00871389 0.560 0.0775174 0.205 -0.0770479 0.00488102 0.469 0.079955 0.00952484 0.409 0.0694818 0.174 -0.0347004 0.708 0.0958863 0.187 -0.0567748 0.641 0.0936098 0.172 -0.0618746 0.632 0.0829943 0.159 -0.0729234 0.00698018 0.482 0.0205063 0.411 -0.0155159 0.166 0.0812484 0.141 -0.0153593 0.262 0.025185 0.686 0.0263891 0.357 0.0598013 0.730 0.185859 0.949 0.0692377 0.257 0.00673472 0.391 -0.0435482 0.050 -0.081526 -0.0273596 -0.0600487 0.0533032 -0.145955 p 0.747 0.864 0.839 0.748 0.734 0.799 0.871 0.598 0.600 0.748 0.129 0.819 0.466 Parcellation Craddock130 Craddock260 Craddock500 Craddock950 Shen100 Shen200 Shen300 OASIS Power264 Talairach Stanford AAL Gordon_surface 0.0381087 Marginally significant individual differentiating FC changes in PD Despite non-reproducibility of PD-related global FC changes across different datasets, a small number of ROI-pairs that distinguish PD from controls may nevertheless, in principle, show reproducible changes across datasets. To check for this possibility, we concentrated on individual brain region pairs with FC changes that are significant w.r.t. all datasets, by sorting the ROI-pairs according to their least significance max(p) with respect to all datasets. For example, Table 6 shows the ROI-pairs with FC decreases in PD (i.e. positive t-values, corresponding to NC>PD) and max(p) < 0.05 for the Power264 parcellation, without global signal regression. The best ROI-pair has max(p+)=0.0125, so min(max(p+))=0.0125. To check whether this min(max(p+)) is statistically significant, we performed permutation tests as described. Table 7 lists these min(max(p)) values as well as their associated significance p(min(max(p))) for all 13 parcellations. Only two out of the 13 parcellations yielded significant ROI-pairs at the p<0.05 significance level ('Power264' and 'Talairach'), while a third parcellation produced only marginally significant ROI-pairs ('Shen100', p=0.055) - see Table 8 and Fig 5. These (marginally) significant ROI-pairs involve visual-sensorimotor, respectively visual-parietal association areas. Whether these ROI-pair changes are more widely reproducible or not will have to await the release of more publicly-available PD rs-fMRI datasets. 19 Table 6. Best ROI-pairs with FC decreases in PD (t>0, corresponding to NC>PD) and max(p) < 0.05 (for the Power264 parcellation, without global signal regression). max(p) ROI1-ROI2 NEUROCON (NC-PD) p p TaoWu (NC- p PPMI (NC- PD) PD) t NEUROCON (NC-PD) t TaoWu (NC-PD) t PPMI (NC- PD) 0.012526 0.016354 0.025311 0.038357 0.038529 0.038634 0.044406 0.046861 sphere5(-21,-31,61)- sphere5(15,-77,31) sphere5(-42,45,-2)- sphere5(43,-78,-12) sphere5(-21,-31,61)- sphere5(-24,-91,19) sphere5(-38,-33,17)- sphere5(20,-86,-2) sphere5(37,-81,1)- sphere5(38,-17,45) sphere5(-21,-31,61)- sphere5(-26,-90,3) sphere5(-21,-31,61)- sphere5(29,-77,25) sphere5(-21,-31,61)- sphere5(-40,-88,-6) 0.012526 0.0018915 0.011638 2.5541 3.3398 0.0026347 0.016354 0.0079673 3.1172 2.5153 0.0019939 0.025311 0.025232 3.2025 2.3298 0.038357 0.00041882 0.014314 2.1104 3.8665 0.038529 0.02727 0.038269 2.1123 2.297 0.012064 0.038634 0.033842 2.5863 2.1425 0.0085123 0.027058 0.044406 2.7193 2.3019 0.0077414 0.046861 0.017283 2.7342 2.0557 2.7402 2.905 2.3949 2.653 2.2003 2.2603 2.1243 2.55 Table 7. Significance of min(max(p)) values for all 13 parcellations (no global signal regression). min(max(p+)) min(max(p)) p(min(max(p+))) p(min(max(p))) 0.0607935 0.0503968 0.0294203 0.0202304 0.0172931 0.0371887 0.0425589 0.0324906 0.0697703 0.012526 0.0275321 0.00701726 0.0599084 0.112189 0.242712 0.148728 0.0909226 0.0542509 0.376232 0.151199 0.115015 0.3402 0.118716 0.11095 0.0339528 0.281706 0.208 0.152 0.170 0.206 0.316 0.055* 0.195 0.199 0.208 0.033* 0.185 0.032* 0.145 0.426 0.831 0.843 0.859 0.815 0.905 0.747 0.804 0.938 0.846 0.846 0.522 0.820 Parcellation AAL Craddock130 Craddock260 Craddock500 Craddock950 Shen100 Shen200 Shen300 OASIS Power264 Gordon_surface Talairach Stanford 20 Table 8. Marginally significant FC changes w.r.t. all 3 datasets (decreased in PD). Parcellation p(min(max(p+))) min(max(p+)) ROI1 Talairach 0.032 0.00701726 Power264 0.033 0.012526 Shen100 0.055 0.0371887 (-24,-58,4) left visual association area, lingual gyrus, BA18 sphere5(-21,-31,61) left postcentral / precentral gyrus L.BA19.3 left cuneus, precuneus ROI2 (–38, –32,16) left superior temporal gyrus, BA41, planum temporale / parietal operculum sphere5(15, –77,31) right cuneus R.BA6.1 right SMA, middle cingulate Fig 5. Marginally significant FC changes w.r.t. all 3 datasets. The ROIs were mapped onto the brain surface using BrainNet Viewer [Xia et al., 2013] (http://www.nitrc.org/projects/bnv/). Testing classifiers for discriminating PD-related FC changes Training classifiers on functional connectivity data for each one of the 3 datasets (NEUROCON, Tao Wu, PPMI) and testing them on the other two datasets produced average accuracies on test data in the range 0.225 - 0.7 (mean 0.497, standard deviation 0.073), while a random classifier is expected to have an average accuracy of 0.5. Fig 6 shows the corresponding average accuracies Aacc = (acc(NC)+acc(PD))/2 for standard preprocessing (with the default FSL highpass filter at 0.01Hz), global signal regression and respectively bandpass filtering (0.01-0.1Hz) for both linear SVM and Gaussian Naïve Bayes (GNB) classifiers with N=5000 features (out of the total of 6670 ROI pairs of the AAL parcellation). See Supplementary Fig 2 and Supplementary Table 5 for the accuracies of classifiers with N=10,50,100,500,5000 features. 21 Since for each training dataset (for example NEUROCON), we have two different test datasets (PPMI and TaoWu in our example), we also computed an aggregated average accuracy for each dataset by taking the mean of the two average accuracies corresponding to the two remaining test datasets (Aacc(dataset-dataset1)+Aacc(dataset-dataset2))/2. The resulting aggregated average accuracies were low, in the range 0.336 - 0.591 (mean 0.497, standard deviation 0.0522, compared to 0.5 for a random classifier; see also Fig 7). Since the three datasets are not technically homogeneous, we also trained and tested classifiers on random splits of the same dataset, to check to what extent the low accuracies are due to technical differences, or to disease heterogeneity. Fig 8 shows the average accuracies for 10,000 random splits in half of each dataset and various preprocessing options. Again, the means of the average accuracies over the 10,000 tests were low, in the range 0.51 – 0.66, reinforcing the evidence for disease heterogeneity. Fig 6. Average accuracies for classifiers trained on dataset 1 and tested on dataset 2 for all dataset pairs using standard preprocessing ('standard'), global signal regression (GS) and respectively bandpass filtering (0.01-0.1Hz). Here, SVM (linear Support Vector Machine) and GNB (Gaussian Naïve Bayes) classifiers used N=5000 features – see Supplementary Fig 2 for classifier accuracies for varying N. As an example, NEUROCON-PPMI denotes classifiers trained on NEUROCON and tested on PPMI data. 22 Fig 7. Aggregated average accuracies for classifiers trained on each of the 3 datasets using standard preprocessing ('standard'), global signal regression (GS) and respectively bandpass filtering (0.01-0.1Hz). Classifiers were trained with N=10,50,100,500,5000 features. As an example, NEUROCON(10) refers to the aggregated accuracy (Aacc(NEUROCON-PPMI) + Aacc(NEUROCON-TaoWu))/2 for classifiers trained on NEUROCON and tested on PPMI and respectively TaoWu data using N=10 features. SVM – linear SVM classifier, GNB – Gaussian Naïve Bayes classifier. Fig 8. Average accuracies for classifiers trained and tested on split data from the same dataset using standard preprocessing ('standard'), global signal regression (GS) and respectively bandpass filtering (0.01-0.1Hz). An SVM classifier with N=5000 features was used. 23 Discussion The accelerated increase in the number of functional connectivity studies of Parkinson's Disease requires a consolidation of the knowledge in this field for enabling the development of clinically relevant rs-fMRI markers for this disease. Unfortunately however, there are many inconsistencies between published works and virtually no high confidence reproducibility studies. This is the first study investigating the reproducibility of functional connectivity changes in Parkinson's disease on more than two datasets. The fact that we use a uniform data processing workflow for all datasets excludes a large number of technical factors as potential culprits for the observed differences between datasets. Also, the fact that our comparison involves three datasets is essential, as it lowers the probability that the observed lack of reproducibility is due to a problematic dataset – in such a case, with 3 datasets we might still observe reproducibility with respect to the remaining dataset pair, something which we do not see in reality. To better clarify the issue, we devised a method to directly check for disease heterogeneity using random splits of a single dataset with replicate scans. Technical differences can then be excluded since all the scans have been acquired under identical technical conditions. The fact that we still observe non-reproducibility in a significant fraction of random subsamples of each individual dataset (these subsamples being technically homogeneous as they come from the same dataset), suggests that functional heterogeneity may be a dominating factor behind the lack of reproducibility of functional connectivity alterations in different resting state fMRI studies of Parkinson's disease. This could be due to the heterogeneous multi-lesional topography and progression of the neurodegenerative process, possibly accompanied by variable compensatory functional circuit changes, as well as by changes due to dopaminergic medication [Tahmasian et al., 2015]. For a a more direct graphical depiction of the heterogeneity of the functional connectomes of the PD patients, we have applied consensus NMF clustering [Brunet et al., 2004] for a progressively increasing number of clusters (k=2,…,18, Supplementary Fig 3). Note that besides the consistent grouping of the replicate scan pairs for each patient, it is difficult to single out an optimal number of clusters k. While global PD-related functional connectivity differences were non-reproducible across datasets, we identified a few individual ROI pairs with marginally consistent FC differences across all three datasets. However, finding out whether these differences are more widely reproducible or not will have to await the release of more public PD datasets. Additionally, we applied more sophisticated multivariate machine learning techniques to learn classifiers that discriminate PD from controls using functional connectivities between ROI pairs as features. However, training classifiers on each one of the three datasets (NEUROCON, Tao Wu, PPMI) produced only low accuracies on the remaining two (test) datasets, in line with the preceding results. Furthermore, since the three datasets are not technically homogeneous, we also trained and tested classifiers on random splits of the same dataset, to more directly check to what extent the low accuracies are due to technical differences, or to disease heterogeneity. Again, we obtained low average accuracies (with means in the range 0.51 – 0.66), reinforcing the evidence for disease heterogeneity. Interestingly, these results are consistent with a recent study [Orban et al., 2017] on multisite generalizability of schizophrenia diagnosis based on functional brain connectivity, which reported multisite classification accuracies below 70%, in contrast to over 30 24 previously published, classification accuracy exceeds 80%. largely single-site schizophrenia studies, whose average reported Therefore, given the paucity of publicly available rs-fMRI PD datasets, we advocate the critical importance of data sharing for enabling the discovery of reproducible and clinically useful functional imaging biomarkers of PD. In this regard, we view our study as an important first step towards more refined reproducibility studies that would be possible only with more publicly available datasets. In view of the many inconsistencies found in the published literature on PD- related functional connectivity changes, we strongly argue for a direct computational comparison of PD rs-fMRI datasets using a uniform data processing workflow, to avoid publication bias as well as processing workflow differences in the separate studies. Limitations. The present study has concentrated on PD-related changes in functional connectivity (loosely viewed as correlations between different regions of interest), rather than changes in fluctuations of the amplitude of the rs-fMRI signal. In a complementary study, [Wu et al., 2015] observed PD-related changes in ALFF, but with rather limited reproducibility. An in-depth analysis of the reproducibility of PD-related differences in the amplitude of fluctuations is out of the scope of the present paper. Acknowledgments We are very grateful to the PPMI consortium for granting us access to the Parkinson's Progression Markers Initiative (PPMI) imaging data. Data used in the preparation of this article were obtained from the PPMI database (www.ppmi-info.org/data). For up-to-date information on the study, visit www.ppmi-info.org. PPMI – a public-private partnership – is funded by the Michael J. Fox Foundation including AbbVie, Avid Radiopharmaceuticals, Biogen, Bristol-Myers Squibb, Covance, GE Healthcare, Genentech, GlaxoSmithKline, Eli Lilly and Company, Lundbeck, Merck, Meso Scale Discovery, Pfizer, Piramal Imaging, Roche, Servier, UCB (Union Chimique Belge) and GOLUB Capital. for Parkinson's Research and funding partners, We also thank Madalina Tivarus and the reviewers for some very useful suggestions for improving the manuscript. Compliance with Ethical Standards Ethical approval: The study was carried out in accordance with The Code of Ethics of the World Medical Association (Declaration of Helsinki) for experiments involving humans and was approved by the ethics committee of the University Emergency Hospital Bucharest. Informed consent: Informed consent was obtained in written form from all individual participants included in the study. 25 References Baggio HC, Segura B, Sala‐Llonch R, Marti MJ, Valldeoriola F, Compta Y, Tolosa E, Junque C. Cognitive impairment and resting‐state network connectivity in Parkinson's disease. Human brain mapping. 2015 Jan 1;36(1):199-212. Baudrexel S, Witte T, Seifried C, von Wegner F, Beissner F, Klein JC, Steinmetz H, Deichmann R, Roeper J, Hilker R. Resting state fMRI reveals increased subthalamic nucleus–motor cortex connectivity in Parkinson's disease. Neuroimage. 2011 Apr 15;55(4):1728-38. Brunet JP, Tamayo P, Golub TR, Mesirov JP. Metagenes and molecular pattern discovery using the national academy of sciences. 2004 Mar matrix factorization. Proceedings of 23;101(12):4164-9. Chen Y, Yang W, Long J, Zhang Y, Feng J, Li Y, Huang B. Discriminative Analysis of Parkinson's Disease Based on Whole-Brain Functional Connectivity. PloS one. 2015 Apr 17;10(4):e0124153. Craddock RC, James GA, Holtzheimer PE, Hu XP, Mayberg HS. A whole brain fMRI atlas generated via spatially constrained spectral clustering. Human brain mapping. 2012 Aug 1;33(8):1914-28. Esposito F, Tessitore A, Giordano A, De Micco R, Paccone A, Conforti R, Pignataro G, Annunziato L, Tedeschi G. Rhythm-specific modulation of the sensorimotor network in drug- naive patients with Parkinson's disease by levodopa. Brain. 2013 Mar 1;136(3):710-25. Fahn S. Recent developments in Parkinson's disease. Edited by S. Fahn, C. D. Mardsen, P. Jenner, and P. Teychenne. Raven Press; 1986. Friston KJ, Buchel C. Functional connectivity. Human Brain Function. 2003;2:999-1018. Gordon EM, Laumann TO, Adeyemo B, Huckins JF, Kelley WM, Petersen SE. Generation and evaluation of a cortical area parcellation from resting-state correlations. Cerebral cortex. 2014 Oct 14:bhu239. Göttlich M, Münte TF, Heldmann M, Kasten M, Hagenah J, Krämer UM. Altered resting state brain networks in Parkinson's disease. PloS one. 2013 Oct 28;8(10):e77336. Griffanti L, Salimi-Khorshidi G, Beckmann CF, Auerbach EJ, Douaud G, Sexton CE, Zsoldos E, Ebmeier KP, Filippini N, Mackay CE, Moeller S. ICA-based artefact removal and accelerated fMRI acquisition for improved resting state network imaging. Neuroimage. 2014 Jul 15;95:232- 47. Hacker C, Perlmutter J, Criswell S, Ances B, Snyder A. Resting state functional connectivity of the striatum in Parkinson's disease. Brain. 2012 Nov 28:aws281. Hayasaka S. Functional connectivity networks with and without global signal correction. Frontiers in human neuroscience 7. 2013. Helmich RC, Derikx LC, Bakker M, Scheeringa R, Bloem BR, Toni I. Spatial remapping of cortico-striatal connectivity in Parkinson's disease. Cerebral cortex. 2010 May 1;20(5):1175-86. Helmich RC, Janssen MJ, Oyen WJ, Bloem BR, Toni I. Pallidal dysfunction drives a cerebellothalamic circuit into Parkinson tremor. Annals of neurology. 2011 Feb 1;69(2):269-81. 26 Hoehn MM, Yahr MD. Parkinsonism: Onset, progression and mortality. Neurology 57:S11–S26. 2001. Kurani AS, Seidler RD, Burciu RG, Comella CL, Corcos DM, Okun MS, MacKinnon CD, Vaillancourt DE. Subthalamic nucleus-sensorimotor cortex functional connectivity in de novo and moderate Parkinson's disease. Neurobiology of aging. 2015 Jan 31;36(1):462-9. Kwak Y, Peltier S, Bohnen N, Müller M, Dayalu P, Seidler RD. Altered resting state cortico- striatal connectivity in mild to moderate stage Parkinson's disease. Frontiers in systems neuroscience. 2010 Sep 15;4:143. Kwak Y, Peltier S, Bohnen N, Müller M, Dayalu P, Seidler RD. L-DOPA changes spontaneous low-frequency BOLD signal oscillations in Parkinson's disease: a resting state fMRI study. Frontiers in systems neuroscience. 2012 Jul 4;6:52. Liu D, Dong Z, Zuo X, Wang J, Zang Y. Eyes-open/eyes-closed dataset sharing for reproducibility evaluation of resting state fMRI data analysis methods. Neuroinformatics. 2013 Oct 1;11(4):469-76. Liu H, Edmiston EK, Fan G, Xu K, Zhao B, Shang X, Wang F. Altered resting-state functional connectivity of the dentate nucleus in Parkinson's disease. Psychiatry Research: Neuroimaging. 2013 Jan 30;211(1):64-71. Long D, Wang J, Xuan M, Gu Q, Xu X, Kong D, Zhang M. Automatic classification of early Parkinson's disease with multi-modal MR imaging. PloS one. 2012 Nov 9;7(11):e47714. Luo C, Song W, Chen Q, Zheng Z, Chen K, Cao B, Yang J, Li J, Huang X, Gong Q, Shang HF. Reduced functional connectivity in early-stage drug-naive Parkinson's disease: a resting-state fMRI study. Neurobiology of aging. 2014 Feb 28;35(2):431-41. Marcus DS, Wang TH, Parker J, Csernansky JG, Morris JC, Buckner RL. Open Access Series of Imaging Studies (OASIS): cross-sectional MRI data in young, middle aged, nondemented, and demented older adults. Journal of cognitive neuroscience. 2007 Sep;19(9):1498-507. Murphy K, Birn RM, Handwerker DA, Jones TB, Bandettini PA. The impact of global signal regression on resting state correlations: are anti-correlated networks introduced?. Neuroimage. 2009 Feb 1;44(3):893-905. Orban P, Dansereau C, Desbois L, Mongeau-Pérusse V, Giguère CÉ, Nguyen H, Mendrek A, Stip E, Bellec P. Multisite generalizability of schizophrenia diagnosis classification based on functional brain connectivity. Schizophrenia research. 2017 Jun 7. Power JD, Cohen AL, Nelson SM, Wig GS, Barnes KA, Church JA, Vogel AC, Laumann TO, Miezin FM, Schlaggar BL, Petersen SE. Functional network organization of the human brain. Neuron. 2011 Nov 17;72(4):665-78. Power JD, Schlaggar BL, Petersen SE. Recent progress and outstanding issues in motion correction in resting state fMRI. Neuroimage. 2015 Jan 15;105:536-51. Salimi-Khorshidi G, Douaud G, Beckmann CF, Glasser MF, Griffanti L, Smith SM. Automatic denoising of functional MRI data: combining independent component analysis and hierarchical fusion of classifiers. Neuroimage. 2014 Apr 15;90:449-68. 27 Sharman M, Valabregue R, Perlbarg V, Marrakchi‐Kacem L, Vidailhet M, Benali H, Brice A, Lehéricy S. Parkinson's disease patients show reduced cortical‐subcortical sensorimotor connectivity. Movement Disorders. 2013 Apr 1;28(4):447-54. Shen X, Tokoglu F, Papademetris X, Constable RT. Groupwise whole-brain parcellation from resting-state fMRI data for network node identification. Neuroimage. 2013 Nov 15;82:403-15. Shirer WR, Ryali S, Rykhlevskaia E, Menon V, Greicius MD. Decoding subject-driven cognitive states with whole-brain connectivity patterns. Cerebral cortex. 2012 Jan 1;22(1):158-65. Skidmore FM, Yang M, Baxter L, Von Deneen KM, Collingwood J, He G, White K, Korenkevych D, Savenkov A, Heilman KM, Gold M. Reliability analysis of the resting state can sensitively and specifically identify the presence of Parkinson disease. Neuroimage. 2013 Jul 15;75:249-61. Szewczyk-Krolikowski K, Menke RA, Rolinski M, Duff E, Salimi-Khorshidi G, Filippini N, Zamboni G, Hu MT, Mackay CE. Functional connectivity in the basal ganglia network differentiates PD patients from controls. Neurology. 2014 Jul 15;83(3):208-14. Talairach J, Tournoux P. Co-planar stereotaxic atlas of the human brain. 3-Dimensional proportional system: an approach to cerebral imaging. 1988 Tessitore A, Amboni M, Esposito F, Russo A, Picillo M, Marcuccio L, Pellecchia MT, Vitale C, Cirillo M, Tedeschi G, Barone P. Resting-state brain connectivity in patients with Parkinson's disease and freezing of gait. Parkinsonism & related disorders. 2012 Jul 31;18(6):781-7. Tuite PJ, Mangia S, Michaeli S. Magnetic resonance imaging (MRI) in Parkinson's disease. Journal of Alzheimer's disease & Parkinsonism. 2013 Mar 25:001. Tzourio-Mazoyer N, Landeau B, Papathanassiou D, Crivello F, Etard O, Delcroix N, Mazoyer B, Joliot M. Automated anatomical labeling of activations in SPM using a macroscopic anatomical parcellation of the MNI MRI single-subject brain. Neuroimage. 2002 Jan 31;15(1):273-89. Wen X, Wu X, Liu J, Li K, Yao L. Abnormal baseline brain activity in non-depressed Parkinson's disease and depressed Parkinson's disease: a resting-state functional magnetic resonance imaging study. PloS one. 2013 May 22;8(5):e63691. Wu T, Long X, Wang L, Hallett M, Zang Y, Li K, Chan P. Functional connectivity of cortical motor areas in the resting state in Parkinson's disease. Human brain mapping. 2011 Sep 1;32(9):1443-57. Wu T, Ma Y, Zheng Z, Peng S, Wu X, Eidelberg D, Chan P. Parkinson's Disease-Related Spatial Covariance Pattern Identified with Resting-State Functional MRI. Journal of Cerebral Blood Flow & Metabolism. 2015 Nov 1;35(11):1764-70. Wu T, Wang J, Wang C, Hallett M, Zang Y, Wu X, Chan P. Basal ganglia circuits changes in Parkinson's disease patients. Neuroscience letters. 2012 Aug 22;524(1):55-9. Xia M, Wang J, He Y. BrainNet Viewer: a network visualization tool for human brain connectomics. PloS one. 2013 Jul 4;8(7):e68910. Yu R, Liu B, Wang L, Chen J, Liu X. Enhanced functional connectivity between putamen and supplementary motor area in Parkinson's disease patients. PLoS One. 2013 Mar 26;8(3):e59717. 28 Zhang D, Liu X, Chen J, Liu B, Wang J. Widespread increase of functional connectivity in Parkinson's disease with tremor: a resting-state FMRI study. Frontiers in aging neuroscience. 2015;7. 29 Supporting Information Resting-state fMRI studies of Parkinson's disease Supplementary Table 1. Resting-state fMRI studies of Parkinson's disease Study Main study characteristics eyes open / closed NC PD Patient characteristics Gottlich 2013 graphs, AAL, AAL subdivision in 343 cortical & subcortical ROIs eyes closed 20 37 advanced disease, on medication (L-dopa, agonists) classifier: SVM (ALFF, ReHo, RFCS, GM, WM, CSF) accuracy 86.96%, sensitivity 78.95%, specificity 92.59%, precision 88% classifier (ALFF): accuracy 88%, sensitivity 92%, specificity 87%, precision 87% eyes closed 27 19 early stage (H&Y 1-2), OFF (12 h) eyes closed 15 14 OFF(12-18 h) Long 2012 Skidmore 2013 Baudrexel 2011 FC STN, M1 hand area eyes closed 44 31 Wu 2011 FC pre-SMA - M1 seeds: pre-SMA, M1, PCC eyes closed 18 18 early stage PD patients (n=31) during the medication-off state with healthy controls (n=44); 16 tremor, 15 non-tremor akinesia right (at most mild tremor), OFF (12h) H&Y 1.78+- 0.5 Kwak 2010 6 striatal seeds: 3 caudate (inferior ventral striatum, superior ventral striatum, dorsal caudate), 3 putamen (dorsal caudal putamen, dorsal rostral putamen, ventral rostral putamen) eyes fixed on cross 24 25 mild to moderate stage (H&Y 1- 2.5), ON and OFF(12–18 h) Kwak 2012 ALFF (fALLF) 24 24 Helmich 2010 FC anterior, posterior putamen, caudate posterior cingulate (control) eyes closed 36 41 Helmich 2011 tremor eyes closed 36 41 mild to moderate stage (H&Y 1- 2.5) ON and OFF (12-18h) 13 PD without any tremor, 18 moderate to severe. 10 - never anti-PD medication (median H&Y 2.1, max 5) perfectly age-matched (57 years) 19 tremor, 23 nontremor, right- handed, 12 no medication, OFF (12 h) Supporting information 30 Luo 2014 FC anterior, posterior putamen, anterior caudate, amygdala eyes closed 52 52 Yu 2013 FC putamen, caudate, and SMA 20 19 52 PD right handed, early stage drug-naïve, H&Y 1.85 (max 3), 31 right onset, 21 left onset OFF obvious at least a mild tremor Hacker 2012 FC striatum 6 ROIs: (i) caudate nucleus; (ii) anterior putamen; and (iii) posterior putamen Kurani 2015 FC (restricted to motor areas) seeds: STN (most affected part), posterior cingulate (control seed) eyes open (fixate cross) eyes open (focused on the word "RELAX") 19 13 advanced PD ON 19 39 20 de novo 19 moderate OFF Baggio 2015 cognitive impairement ICA+dual regression (25 networks), group comparison of: DMN, dorsal attention network (DAN), bilateral frontoparietal networks (FPN), 43 seeds (10 DAN, 18 DMN, 15 FPN) 36 65 ON state 65 nondemented: 34% mild cognitive impairement (MCI) Esposito 2013 ICA (fastICA BrainVoyager 40 components, restriction to sensorimotor network-best fit with previous template) eyes closed 18 20 20 drug naïve PD, 10 before & after (1H) levodopa, 10 before & after placebo Szewczyk- Krolikowski 2014 ICA+dual regression – basal ganglia (BG) network from 80 separate controls - classifier based on average BG component values in voxels discriminating PD(OFF)-controls eyes open 19 19 discovery cohort: 19 PD ON/OFF, 19 controls validation cohort: 13 PD (5 drug- naïve), no controls (this disallows evaluation of specificity!) 80 elderly controls for BG template (MELODIC groupICA 50 components) all subjects right-handed Tessitore 2012 fastICA, sogICA 40 components select DMN, the frontoparietal (right and left FPN), sensorimotor network (SMN), visual, auditory eyes closed 15 29 PD ON 16 FOG+ Supporting information 31 ROIs: caudate, putamen, globus pallidus, thalamus sensorimotor (M1, postcentral gyrus) associative (ventrolateral, dorsolateral prefrontal) limbic (orbitofrontal, rectus gyrus, cingulate, insula, medial temporal ctx, perirhinal & entorhinal cortex, hippocampus, amygdala) ROIs: SNc bilaterally effective connectivity Granger causality analysis (GCA) with long TR=2000 & short TR=400 Sharman 2013 Wu 2012 eyes closed 45 36 eyes closed 16 16 16 de novo OFF/ON Liu 2013 FC dentate nucleus (cerebellum) -> cerebellar output eyes closed 18 18 Chen 2015 SVM classifier LOOCV based on FC between 116 ROIs of AAL parcellation (150 features selected by Kendall tau correlation) - 93.62% accuracy, 90.47% sensitivity, 96.15 specificity eyes closed 26 21 mild to moderate (1.34 H&Y average) in OFF state 8 rigidity & bradykinesia- dominant (PD_AR) 10 tremor-dominant (PD_T) 21 PD (10 males, 11 females, 58.3 years), 26 HC (10 males, 16 females, 61.3 years) OFF state (12 hours) UPDRS 29.8 (sd 9.1), disease duration 3.2 years (sd 3.2) Wen 2013 depression ALFF eyes closed 21 33 OFF 17 depression Supporting information 32 Subject motion in scanner Since subject motion in the scanner has been observed to have significant influence on the functional connectivities computed from rs-fMRI data, despite motion correction (e.g. [Power et al., 2015]), we also considered subsets of scans with low in-scanner motion (marked by the suffix '0', e.g. 'NC0' and 'PD0' – see also Supplementary Table 2). The 3 datasets have different in-scanner motion characteristics. Tao Wu Low motion condition Supplementary Table 2. The subsets of scans with low in-scanner motion Dataset NEUROCON max abs motion  1/4 voxel = 0.958mm max rel motion  1/6 voxel = 0.638mm max abs motion  1/4 voxel = 1mm max rel motion  1/6 voxel = 0.67mm max abs motion  1/3 voxel = 1.098mm max rel motion  1/3 voxel = 1.098mm mean abs motion  1/6 voxel = 0.549mm max rel motion  1/6 voxel = 0.549mm PPMI 134 20 NC scans PD scans NC0 scans PD0 scans 31 39 54 27 20 19 13 13 14 89 The following Tables show the p-values corresponding to potential group differences in motion in the 3 datasets analyzed: NEUROCON NC-PD (31-54) p-value (t-value) p(mean rel) p(mean abs) p(max rel) p(max abs) 0.00243021 (-3.14036) 0.0545481 (-1.96039) 0.0945346 (-1.6961) 0.135713 (-1.50937) NC0-PD0 (27-39) p(mean rel) p(mean abs) p(max rel) p(max abs) 0.166962 (-1.39829) 0.043621 (-2.06138) 0.319776 (-1.00337) 0.100148 (-1.67026) Note that in the NEUROCON cohort, patients move more than normal controls. Supporting information 33 Tao Wu NC-PD (20-20) p(mean rel) p(mean abs) p(max rel) p(max abs) 0.0270361 (2.33514) 0.727106 (0.352521) 0.0074933 (2.96994) 0.543508 (0.613522) NC0-PD0 (13-14) p(mean rel) p(mean abs) p(max rel) p(max abs) 0.256895 (1.16397) 0.497445 (0.688602) 0.0859714 (1.80121) 0.455533 (0.758043) In the Tao Wu cohort, patients move less than normal controls. PPMI NC-PD (19-134) p(mean rel) p(mean abs) p(max rel) p(max abs) 0.0619124 (1.9849) 0.485967 (0.710965) 0.107866 (1.69169) 0.134632 (1.56558) NC0-PD0 (13-89) p(mean rel) p(mean abs) p(max rel) p(max abs) 0.0656489 (1.94269) 0.314195 (-1.03452) 0.224242 (1.26499) 0.53616 (0.63367) NC_center32-PD_center32 (9-30) p(mean rel) p(mean abs) p(max rel) p(max abs) 0.242348 (1.23728) 0.77672 (0.289486) 0.344483 (1.00257) 0.318376 (1.05993) NC0_center32-PD0_center32 (7-23) p(mean rel) p(mean abs) p(max rel) p(max abs) 0.173686 (1.40637) 0.649951 (0.46271) 0.252684 (1.21065) 0.13755 (1.64485) Supporting information 34 ROI-pairs with significant group differences in the separate datasets Supplementary Table 3 below shows the numbers of significant ROI pairs for several significance thresholds for the unpaired t-test between patient and control functional connectivities (for the AAL parcellation and without correction for multiple comparisons, due to the limited sample sizes). A single scan for each subject was considered in this comparison. Supplementary Table 3. Number of significant ROI-pairs for the AAL parcellation significance NEUROCON(16-27) TaoWu(20-20) PPMI(18-91) level 0.001 0.005 0.01 0.05 4 23 48 283 15 68 121 489 19 81 156 737 As expected, the power of the tests was greater for the datasets with more samples, but the effect sizes (t-values) were similar, as can be seen in Supplementary Table 4 below, which shows the top positive and respectively negative effect sizes observed. Supplementary Table 4. Top effect sizes (t-values) and the corresponding p-values for the top positive and respectively negative FC alterations top t+ top p+ top t− top p− NEUROCON(16-27) TaoWu(20-20) PPMI(18-91) 3.05647 0.00429499 -4.495779 6.09E-05 4.632511 4.23E-05 -2.91725 0.0062734 4.70764 7.99E-05 -3.870344 0.0003765 Supporting information 35 Comparison of random splits of an 'eyes open-eyes closed' dataset Besides the potentially heterogeneous contrast between PD and normal controls, we also tested our method on a different, potentially more homogeneous contrast, namely 'eyes open' versus 'eyes closed' resting state in healthy volunteers. We used the Beijing eyes-open-eyes closed (EO-EC) dataset [Liu et al., 2013] (http://fcon_1000.projects.nitrc.org/indi/IndiPro.html), which involved 48 college students aged 19–31 years, 24 female with no history of neurological and psychiatric disorders. Each participant underwent three 8 min resting state scanning sessions: an EC session followed by two sessions counter-balanced across subjects: one EO resting state and one EC resting state session. The functional images were obtained on a Siemens Trio 3 Tesla scanner using an echo-planar imaging sequence with the following parameters: 33 axial slices, thickness/gap=3.5/0.7 mm, in- plane resolution=64×64, repetition time=2000 ms, echo time=30 ms, flip angle=90°, field of view (FOV)=200×200mm2, 240 volumes per scan. In addition, a 3D T1-weighted MPRAGE image was acquired with the following parameters: 128 sagittal slices, slice thickness/gap=1.33/0 mm, in- plane resolution=256×192, TR=2530 ms, TE=3.39 ms, inversion time (TI)=1100 ms, flip angle=7°, FOV=256×256 mm2. Note that the parameters used in this study are quite similar to the ones from the PD datasets, including the scanning time (~8 min), with the exception of the 1.5 Tesla field strength used in the NEUROCON study (all the other studies used 3 Tesla machines). We repeated our analyses of reproducibility of group changes in functional connectivity on random splits of the Beijing EO-EC dataset on both "split subjects" (heterogeneous) and "split replicates" (homogeneous) datasets using the AAL parcellation. As in the case of PD, permutation tests were employed to compute p-values of the reproducibility across split datasets. Additionally, we repeated the analysis for the data with global signal regression. 'Eyes Open-Eyes Closed' FC changes are reproducible The fact that the well-known clinical heterogeneity of Parkinson's disease is also accompanied by heterogeneity in resting state functional connectivity may not retrospectively be a big surprise to an experienced neurologist, although its exact extent could not have been estimated a priori, before analyzing the data. Supporting information 36 However, does this FC heterogeneity in PD also imply the lack of practical usefulness of rs-fMRI functional connectivity? Are there any other conditions that can be reliably differentiated using resting state functional connectivity? To answer these questions, we applied our approach to a different, potentially more homogeneous contrast, namely that between eyes open and eyes closed resting state conditions in healthy volunteers. Repeating our analysis of reproducibility of FC group changes on random splits of the Beijing eyes open-eyes closed dataset [Liu et al., 2013] revealed reproducibility (p<0.05) not just in the homogeneous dataset splits, but also in the heterogeneous ones (Supplementary Fig 1 – only 6% of the heterogeneous and just 0.8% of the homogeneous random splits were non-reproducible at the p>0.05 level). This implies that the EO-EC contrast produces more homogeneous and reproducible global FC changes. Supplementary Fig 1. Consistent reproducibility of 'eyes open'-'eyes closed' (EO-EC) FC changes in random heterogeneous and homogeneous dataset splits. (A) Complementary cumulative distribution function (CCDF=1-CDF) of the reproducibility p-values for Ns=1056 random heterogeneous splits and Ns=814 random homogeneous splits. (B) CCDF of the reproducibility measure R. (C) A scatter-plot of ROI-pair t-values for a random heterogeneous split. (D) A scatter-plot of ROI-pair t-values for a random homogeneous split. Supporting information 37 Learning classifiers for discriminating PD-related FC changes Supplementary Fig 2. Average accuracies Aacc = (acc(NC)+acc(PD))/2 for classifiers trained on dataset 1 and tested on dataset 2 for all dataset pairs using standard preprocessing ('standard'), global signal regression (GS) and respectively bandpass filtering (0.01-0.1Hz). Classifiers were trained with N=10,50,100,500,5000 features. For example, NEUROCON- PPMI(10) shows average accuracies of classifiers trained on NEUROCON and tested on PPMI data using N=10 features. SVM – linear SVM classifier, GNB – Gaussian Naïve Bayes classifier. Supporting information 38 Supplementary Table 5. Detailed performance metrics for the various classifiers TRAIN DATASET TEST DATASET classifier N pairs Pre- processing TEST AvgaccPer Class TEST accPerCl ass(NC) TEST accPerCla ss(PD) NEUROCON PPMI gnb_pooled 10 standard 0.526709 0.1579 0.895522 NEUROCON PPMI svm_linear 10 standard 0.452082 0.1579 0.746269 NEUROCON PPMI gnb_pooled 50 standard 0.447761 0 NEUROCON PPMI svm_linear 50 standard 0.462687 0 0.895522 0.925373 NEUROCON PPMI gnb_pooled 100 standard 0.410644 0.05263 0.768657 NEUROCON PPMI svm_linear 100 standard 0.421642 0 0.843284 NEUROCON PPMI gnb_pooled 500 standard 0.433032 0.05263 0.813433 NEUROCON PPMI svm_linear 500 standard 0.49293 0.10526 0.880597 NEUROCON PPMI gnb_pooled 5000 standard 0.466614 0.05263 0.880597 NEUROCON PPMI svm_linear 5000 standard 0.530833 0.26316 0.798507 NEUROCON TaoWu gnb_pooled 10 standard 0.35 NEUROCON TaoWu svm_linear 10 standard 0.4 NEUROCON TaoWu gnb_pooled 50 standard 0.225 NEUROCON TaoWu svm_linear 50 standard 0.5 NEUROCON TaoWu gnb_pooled 100 standard 0.275 NEUROCON TaoWu svm_linear 100 standard 0.45 NEUROCON TaoWu gnb_pooled 500 standard NEUROCON TaoWu svm_linear 500 standard NEUROCON TaoWu gnb_pooled 5000 standard NEUROCON TaoWu svm_linear 5000 standard 0.325 0.425 0.35 0.45 0.2 0.15 0.15 0.15 0.15 0.05 0.2 0.05 0.25 0.1 0.5 0.65 0.3 0.85 0.4 0.85 0.45 0.8 0.45 0.8 PPMI PPMI PPMI PPMI PPMI PPMI PPMI PPMI PPMI PPMI PPMI PPMI PPMI PPMI PPMI PPMI PPMI PPMI PPMI PPMI NEUROCON gnb_pooled 10 standard 0.491637 0.6129 0.37037 NEUROCON svm_linear 10 standard 0.471326 0.3871 0.555556 NEUROCON gnb_pooled 50 standard 0.486858 0.67742 0.296296 NEUROCON svm_linear 50 standard 0.501493 0.35484 0.648148 NEUROCON gnb_pooled 100 standard 0.519116 0.74194 0.296296 NEUROCON svm_linear 100 standard 0.490143 0.25807 0.722222 NEUROCON gnb_pooled 500 standard 0.576762 0.83871 0.314815 NEUROCON svm_linear 500 standard 0.471924 0.12903 0.814815 NEUROCON gnb_pooled 5000 standard 0.542413 0.67742 0.407407 NEUROCON svm_linear 5000 standard 0.439665 0.06452 0.814815 TaoWu TaoWu TaoWu TaoWu TaoWu TaoWu TaoWu TaoWu TaoWu TaoWu gnb_pooled 10 standard 0.45 0.85 svm_linear 10 standard 0.7 gnb_pooled 50 standard 0.475 svm_linear 50 standard 0.55 gnb_pooled 100 standard 0.5 svm_linear 100 standard 0.625 gnb_pooled 500 standard 0.55 svm_linear 500 standard 0.575 gnb_pooled 5000 standard 0.6 svm_linear 5000 standard 0.575 0.6 0.8 0.3 0.85 0.35 0.9 0.15 0.85 0.15 0.05 0.8 0.15 0.8 0.15 0.9 0.2 1 0.35 1 TaoWu TaoWu TaoWu TaoWu NEUROCON gnb_pooled 10 standard 0.397551 0.25807 0.537037 NEUROCON svm_linear 10 standard 0.41368 0.29032 0.537037 NEUROCON gnb_pooled 50 standard 0.462366 0.25807 0.666667 NEUROCON svm_linear 50 standard 0.397252 0.3871 0.407407 Supporting information 39 TaoWu TaoWu TaoWu TaoWu TaoWu TaoWu TaoWu TaoWu TaoWu TaoWu TaoWu TaoWu TaoWu TaoWu TaoWu TaoWu NEUROCON gnb_pooled 100 standard 0.480884 0.25807 0.703704 NEUROCON svm_linear 100 standard 0.443548 0.3871 0.5 NEUROCON gnb_pooled 500 standard 0.503883 0.32258 0.685185 NEUROCON svm_linear 500 standard 0.41368 0.29032 0.537037 NEUROCON gnb_pooled 5000 standard 0.476105 0.32258 0.62963 NEUROCON svm_linear 5000 standard 0.448626 0.19355 0.703704 PPMI PPMI PPMI PPMI PPMI PPMI PPMI PPMI PPMI PPMI gnb_pooled 10 standard 0.493912 0.36842 0.619403 svm_linear 10 standard 0.539081 0.47368 0.604478 gnb_pooled 50 standard 0.629419 0.68421 0.574627 svm_linear 50 standard 0.499018 0.73684 0.261194 gnb_pooled 100 standard 0.614493 0.68421 0.544776 svm_linear 100 standard 0.518068 0.84211 0.19403 gnb_pooled 500 standard 0.599568 0.68421 0.514925 svm_linear 500 standard 0.483896 0.68421 0.283582 gnb_pooled 5000 standard 0.625687 0.68421 0.567164 svm_linear 5000 standard 0.547722 0.78947 0.30597 NEUROCON PPMI gnb_pooled 10 GS 0.515318 0.10526 0.925373 NEUROCON PPMI svm_linear 10 GS 0.458955 0 0.91791 NEUROCON PPMI gnb_pooled 50 GS 0.575805 0.31579 0.835821 NEUROCON PPMI svm_linear 50 GS 0.571485 0.1579 0.985075 NEUROCON PPMI gnb_pooled 100 GS 0.526905 0.21053 0.843284 NEUROCON PPMI svm_linear 100 GS 0.522584 0.05263 0.992537 NEUROCON PPMI gnb_pooled 500 GS 0.49293 0.10526 0.880597 NEUROCON PPMI svm_linear 500 GS 0.500393 0.10526 0.895522 NEUROCON PPMI gnb_pooled 5000 GS 0.522977 0.1579 0.88806 NEUROCON PPMI svm_linear 5000 GS 0.538885 0.42105 0.656716 NEUROCON TaoWu gnb_pooled 10 GS NEUROCON TaoWu svm_linear 10 GS NEUROCON TaoWu gnb_pooled 50 GS NEUROCON TaoWu svm_linear 50 GS NEUROCON TaoWu gnb_pooled 100 GS NEUROCON TaoWu svm_linear 100 GS NEUROCON TaoWu gnb_pooled 500 GS NEUROCON TaoWu svm_linear 500 GS NEUROCON TaoWu gnb_pooled 5000 GS NEUROCON TaoWu svm_linear 5000 GS 0.5 0.425 0.5 0.4 0.475 0.425 0.45 0.6 0.375 0.575 0.2 0.15 0.4 0.05 0.45 0 0.6 0.25 0.55 0.15 0.8 0.7 0.6 0.75 0.5 0.85 0.3 0.95 0.2 1 PPMI PPMI PPMI PPMI PPMI PPMI PPMI PPMI PPMI PPMI PPMI NEUROCON gnb_pooled 10 GS 0.623656 0.58065 0.666667 NEUROCON svm_linear 10 GS 0.499403 0.25807 0.740741 NEUROCON gnb_pooled 50 GS 0.549881 0.45161 0.648148 NEUROCON svm_linear 50 GS 0.488053 0.16129 0.814815 NEUROCON gnb_pooled 100 GS 0.462366 0.25807 0.666667 NEUROCON svm_linear 100 GS 0.51135 0.09677 0.925926 NEUROCON gnb_pooled 500 GS 0.595878 0.58065 0.611111 NEUROCON svm_linear 500 GS 0.502091 0.09677 0.907407 NEUROCON gnb_pooled 5000 GS 0.616487 0.67742 0.555556 NEUROCON svm_linear 5000 GS 0.423536 0.03226 0.814815 TaoWu gnb_pooled 10 GS 0.525 0.45 0.6 Supporting information 40 PPMI PPMI PPMI PPMI PPMI PPMI PPMI PPMI PPMI TaoWu TaoWu TaoWu TaoWu TaoWu TaoWu TaoWu TaoWu TaoWu TaoWu TaoWu TaoWu TaoWu TaoWu TaoWu TaoWu TaoWu TaoWu TaoWu TaoWu TaoWu TaoWu TaoWu TaoWu TaoWu TaoWu TaoWu TaoWu TaoWu svm_linear 10 GS gnb_pooled 50 GS svm_linear 50 GS gnb_pooled 100 GS svm_linear 100 GS gnb_pooled 500 GS svm_linear 500 GS gnb_pooled 5000 GS svm_linear 5000 GS 0.575 0.45 0.575 0.575 0.65 0.475 0.6 0.55 0.575 0.4 0.45 0.3 0.45 0.35 0.35 0.2 0.55 0.2 0.75 0.45 0.85 0.7 0.95 0.6 1 0.55 0.95 NEUROCON gnb_pooled 10 GS 0.459976 0.29032 0.62963 NEUROCON svm_linear 10 GS 0.434588 0.25807 0.611111 NEUROCON gnb_pooled 50 GS 0.413978 0.16129 0.666667 NEUROCON svm_linear 50 GS 0.446237 0.22581 0.666667 NEUROCON gnb_pooled 100 GS 0.434588 0.25807 0.611111 NEUROCON svm_linear 100 GS 0.441458 0.29032 0.592593 NEUROCON gnb_pooled 500 GS 0.436679 0.35484 0.518519 NEUROCON svm_linear 500 GS 0.443548 0.3871 0.5 NEUROCON gnb_pooled 5000 GS 0.408901 0.35484 0.462963 NEUROCON svm_linear 5000 GS 0.452808 0.3871 0.518519 PPMI PPMI PPMI PPMI PPMI PPMI PPMI PPMI PPMI PPMI gnb_pooled 10 GS 0.536332 0.73684 0.335821 svm_linear 10 GS 0.513354 0.57895 0.447761 gnb_pooled 50 GS 0.520228 0.42105 0.619403 svm_linear 50 GS 0.388845 0.21053 0.567164 gnb_pooled 100 GS 0.512962 0.47368 0.552239 svm_linear 100 GS 0.535546 0.52632 0.544776 gnb_pooled 500 GS 0.630204 0.89474 0.365672 svm_linear 500 GS 0.532011 0.57895 0.485075 gnb_pooled 5000 GS 0.577965 0.89474 0.261194 svm_linear 5000 GS 0.562451 0.73684 0.38806 NEUROCON PPMI gnb_pooled 10 bandpass 0.519049 0.10526 0.932836 NEUROCON PPMI svm_linear 10 bandpass 0.437156 0.1579 0.716418 NEUROCON PPMI gnb_pooled 50 bandpass 0.459348 0.10526 0.813433 NEUROCON PPMI svm_linear 50 bandpass 0.462883 0.05263 0.873134 NEUROCON PPMI gnb_pooled 100 bandpass 0.467203 0.21053 0.723881 NEUROCON PPMI svm_linear 100 bandpass 0.440888 0.1579 0.723881 NEUROCON PPMI gnb_pooled 500 bandpass 0.425766 0.10526 0.746269 NEUROCON PPMI svm_linear 500 bandpass 0.481932 0.1579 0.80597 NEUROCON PPMI gnb_pooled 5000 bandpass 0.414375 0.05263 0.776119 NEUROCON PPMI svm_linear 5000 bandpass 0.497054 0.21053 0.783582 NEUROCON TaoWu gnb_pooled 10 bandpass 0.475 NEUROCON TaoWu svm_linear 10 bandpass 0.525 NEUROCON TaoWu gnb_pooled 50 bandpass 0.475 NEUROCON TaoWu svm_linear 50 bandpass 0.5 NEUROCON TaoWu gnb_pooled 100 bandpass 0.525 NEUROCON TaoWu svm_linear 100 bandpass 0.575 NEUROCON TaoWu gnb_pooled 500 bandpass 0.475 NEUROCON TaoWu svm_linear 500 bandpass 0.7 0.15 0.35 0.35 0.25 0.45 0.35 0.35 0.4 0.8 0.7 0.6 0.75 0.6 0.8 0.6 1 Supporting information 41 NEUROCON TaoWu gnb_pooled 5000 bandpass 0.5 NEUROCON TaoWu svm_linear 5000 bandpass 0.65 0.5 0.35 0.5 0.95 PPMI PPMI PPMI PPMI PPMI PPMI PPMI PPMI PPMI PPMI PPMI PPMI PPMI PPMI PPMI PPMI PPMI PPMI PPMI PPMI TaoWu TaoWu TaoWu TaoWu TaoWu TaoWu TaoWu TaoWu TaoWu TaoWu TaoWu TaoWu TaoWu TaoWu TaoWu TaoWu TaoWu TaoWu TaoWu TaoWu NEUROCON gnb_pooled 10 bandpass 0.448626 0.19355 0.703704 NEUROCON svm_linear 10 bandpass 0.464456 0.35484 0.574074 NEUROCON gnb_pooled 50 bandpass 0.469235 0.29032 0.648148 NEUROCON svm_linear 50 bandpass 0.432796 0.03226 0.833333 NEUROCON gnb_pooled 100 bandpass 0.459976 0.29032 0.62963 NEUROCON svm_linear 100 bandpass 0.483572 0.09677 0.87037 NEUROCON gnb_pooled 500 bandpass 0.485364 0.32258 0.648148 NEUROCON svm_linear 500 bandpass 0.488351 0.03226 0.944444 NEUROCON gnb_pooled 5000 bandpass 0.476105 0.32258 0.62963 NEUROCON svm_linear 5000 bandpass 0.481481 0 0.962963 TaoWu TaoWu TaoWu TaoWu TaoWu TaoWu TaoWu TaoWu TaoWu TaoWu gnb_pooled 10 bandpass 0.475 svm_linear 10 bandpass 0.55 gnb_pooled 50 bandpass 0.5 svm_linear 50 bandpass 0.475 gnb_pooled 100 bandpass 0.475 svm_linear 100 bandpass 0.45 gnb_pooled 500 bandpass 0.475 svm_linear 500 bandpass 0.525 gnb_pooled 5000 bandpass svm_linear 5000 bandpass 0.45 0.55 0.2 0.5 0.4 0.15 0.45 0.2 0.4 0.1 0.45 0.1 0.75 0.6 0.6 0.8 0.5 0.7 0.55 0.95 0.45 1 NEUROCON gnb_pooled 10 bandpass 0.543309 0.29032 0.796296 NEUROCON svm_linear 10 bandpass 0.515532 0.29032 0.740741 NEUROCON gnb_pooled 50 bandpass 0.515532 0.29032 0.740741 NEUROCON svm_linear 50 bandpass 0.364994 0.32258 0.407407 NEUROCON gnb_pooled 100 bandpass 0.499403 0.25807 0.740741 NEUROCON svm_linear 100 bandpass 0.41129 0.32258 0.5 NEUROCON gnb_pooled 500 bandpass 0.506272 0.29032 0.722222 NEUROCON svm_linear 500 bandpass 0.506272 0.29032 0.722222 NEUROCON gnb_pooled 5000 bandpass 0.503883 0.32258 0.685185 NEUROCON svm_linear 5000 bandpass 0.490143 0.25807 0.722222 PPMI PPMI PPMI PPMI PPMI PPMI PPMI PPMI PPMI PPMI gnb_pooled 10 bandpass 0.568932 0.47368 0.664179 svm_linear 10 bandpass 0.531422 0.42105 0.641791 gnb_pooled 50 bandpass 0.607031 0.68421 0.529851 svm_linear 50 bandpass 0.480754 0.84211 0.119403 gnb_pooled 100 bandpass 0.621956 0.68421 0.559701 svm_linear 100 bandpass 0.454635 0.84211 0.067164 gnb_pooled 500 bandpass 0.652003 0.73684 0.567164 svm_linear 500 bandpass 0.473095 0.78947 0.156716 gnb_pooled 5000 bandpass 0.633346 0.73684 0.529851 svm_linear 5000 bandpass 0.547722 0.78947 0.30597 Supporting information 42 Consensus NMF clustering For a more direct graphical depiction of the heterogeneity of the functional connectomes of the PD patient scans, we have applied consensus NMF clustering [Brunet et al., 2004] for a progressively increasing number of clusters k=2,…,18 (Supplementary Fig 3). The Figure depicts the symmetric consensus co-clustering matrices for the PD scans from the NEUROCON dataset. Note that besides the consistent grouping of the replicate scan pairs for each patient, it is difficult to single out an optimal number of clusters k. Supplementary Fig 3. Consensus NMF clustering of functional connectomes of PD patient scans from the NEUROCON dataset. References Brunet JP, Tamayo P, Golub TR, Mesirov JP. Metagenes and molecular pattern discovery using matrix factorization. Proceedings of the national academy of sciences. 2004 Mar 23;101(12):4164- 9. Supporting information 43
1606.08313
1
1606
2016-06-27T15:15:35
Integrated Information and Metastability in Systems of Coupled Oscillators
[ "q-bio.NC" ]
It has been shown that sets of oscillators in a modular network can exhibit a rich variety of metastable chimera states, in which synchronisation and desynchronisation coexist. Independently, under the guise of integrated information theory, researchers have attempted to quantify the extent to which a complex dynamical system presents a balance of integrated and segregated activity. In this paper we bring these two areas of research together by showing that the system of oscillators in question exhibits a critical peak of integrated information that coincides with peaks in other measures such as metastability and coalition entropy.
q-bio.NC
q-bio
Integrated Information and Metastability in Systems of Coupled Oscillators Pedro A.M. Mediano,∗ Juan Carlos Farah, and Murray Shanahan Department of Computing, Imperial College London (Dated: June 28, 2016) It has been shown that sets of oscillators in a modular network can exhibit a rich variety of metastable chimera states, in which synchronisation and desynchronisation coexist. Independently, under the guise of integrated information theory, researchers have attempted to quantify the extent to which a complex dynamical system presents a balance of integrated and segregated activity. In this paper we bring these two areas of research together by showing that the system of oscillators in question exhibits a critical peak of integrated information that coincides with peaks in other measures such as metastability and coalition entropy. Keywords: Synchronisation, chimera states, metastability, integrated information I. INTRODUCTION Systems of coupled oscillators are ubiquitous both in nature and in the human-engineered environment, making them of considerable scientific interest [1]. A variety of mathematical models of such systems have been devised and their synchronisation properties have been the sub- ject of much study. Typical studies of this sort, such as the classic work of Kuramoto [2], examine the conditions under which the system converges on a stable state of either full synchronisation or desynchronisation, perhaps identifying an order parameter that determines a critical phase transition from one state to the other. The collec- tion of known attractors of such systems was enlarged with the discovery of so-called chimera states, in which the system of oscillators partitions into two stable subsets, one of which is fully synchronised while the other remains permanently desynchronised [3]. Although systems of coupled oscillators that converge on stable states are both mathematically interesting and scientifically relevant, they are by no means representative of all real-world synchronisation phenomena. For exam- ple, the brain exhibits synchronous rhythmic activity on multiple spatial and temporal scales, but never settles into a stable state. Although it enters chimera-like states of high partial synchronisation, these are only temporary. A system of coupled oscillators that continually moves from one highly synchronised state to another under its own intrinsic dynamics is said to be metastable. In [4], it was shown that a modular network of phase-lagged Kuramoto oscillators will exhibit metastable chimera states under certain conditions. Variants of this model have since been used to replicate the statistics of the brain under a variety of conditions, including the resting state [5], cognitive control [6], and anaesthesia [7]. In a separate line of enquiry, a number of researchers have attempted to pin down the notion of dynamical complexity. A system is said to have high dynamical com- plexity if it exhibits a balance of integrated and segregated activity, where a system's activity is integrated to the ∗ [email protected] extent that its parts influence each other and segregated to the extent that its parts act independently [8]. Promi- nent among these attempts is the Integrated Information Theory (IIT), originally proposed by Balduzzi and Tononi [9], but extended by multiple authors ever since [10–12]. In the present paper, we connect these two lines of en- quiry by demonstrating that modular networks of coupled oscillators of the sort described in [4] not only exhibit metastable chimera states, but also have high dynamical complexity. Moreover, we show that measures of both phenomena peak in the narrow critical region of the pa- rameter space wherein the system is poised between order and disorder, and IIT offers a rich picture of dynamical complexity in this critical regime. To our knowledge, this is the first description of a dynamical system in which the three major complexity indicators of criticality, metasta- bility, and integrated information all appear. II. METHODS We examine a system of coupled Kuramoto oscillators, extensively used to study non-linear dynamics and syn- chronisation processes [13]. We build upon the work of [4] with a community-structured network of oscillators. The network is composed of 8 communities of 32 oscillators each, with every oscillator being coupled to all other os- cillators in its community with probability 1 and to each oscillator in the rest of the network with probability 1/32. The state of each oscillator i is captured by its phase θi, the evolution of which is governed by the equation Kij sin (θj − θi − α) , (cid:88) = ω + (1) dθi dt 1 κ + 1 j where ω is the natural frequency of each oscillator, κ is the average degree of the network, K is the connectivity matrix and α is a global phase lag. We set ω = 1 and κ = 63. To reflect the community structure, the coupling between two oscillators i, j is Kij = 0.6 if they are in the same community or Kij = 0.4 otherwise. We tune the system by modifying the value of the phase lag, parametrised by β = π/2 − α. We note that the system is fully deterministic, i.e. there is no noise injected in the dynamical equations. A. Metastability In this section we review the basic concepts behind metastability and how it is quantified, following a similar description to that of [4]. The building block of the dynamical quantities we study in this article is the instantaneous synchronisation R, that quantifies the dispersion in θ-space of a given set of oscil- lators. In general, we denote as Rc(t) the instantaneous synchronisation of a community c of oscillators at time t, given by Rc(t) = (cid:104)eiθj (t)(cid:105)j∈c . (2) To quantify metastability, we use the metastability index λ, which is defined as the average temporal variance of the synchrony of each community c, i.e. λc = vart Rc(t) λ = (cid:104)λc(cid:105)c . (3a) (3b) Last, we also define global synchrony ξ as the average across time and space of instantaneous synchrony, ξ =(cid:12)(cid:12)(cid:10)Rc(t)(cid:11) (cid:12)(cid:12) . t,c (4) 2 as a whole compared to when it is considered as the sum of two subsystems M{1,2}. We refer to τ as the integration timescale. In other words, ϕ evaluates how much information is generated by the system but not by the two subsystems alone. For a specific bipartition B = {M 1, M 2}, the effective information of the system beyond B is ϕ[X; τ,B] = I(Xt−τ , Xt) − 2(cid:88) I(M k t−τ , M k t ) . (5) k=1 The main idea behind the computation of Φ is to ex- haustively search all possible partitions of the system and calculate the effective information of each of them. Among those we select the partition with lowest ϕ (under some considerations, see below), termed the Minimum Informa- tion Bipartition (MIB). Then, the integrated information of the system is the effective information beyond its MIB. Given the above expression for ϕ, Φ is defined as Φ[X, τ ] = ϕ[X; τ,BMIB] ϕ[X; τ,B] BMIB = argB min K(B) = min(cid:8)H(M 1), H(M 2)(cid:9) , K(B) (6a) (6b) (6c) According to Eq. (2), Rc (and therefore ξ) is bounded in the [0, 1] interval. Rc(t) will be 1 if all oscillators in c have the same phase at time t, and will be 0 if they are maximally spread across the unit circle. This [0, 1] bound on R allows us to place an upper bound on λ – assuming a unimodal synchrony distribution, the maximum possible value of λ is λmax = 1/9. As defined in Eq. (3a), λc represents the size of the fluctuations in the internal synchrony of a community. A system that is either hypersynchronised or completely desynchronised will have a very small λc, whereas one whose elements fluctuate in and out of synchrony will have a high λc. In other words, a system of oscillators exhibits metastability if its elements remain in the vicinity of a synchronised state without falling into such a state permanently. B. Integrated Information Although other information-theoretic quantities have also been linked to complexity in a neuroscience context [7, 14], we take integrated information Φ as the main informational measure of complexity in our study [9]. There are more modern accounts of the theory [15, 16], but the latest versions have not been as thoroughly studied and are not amenable to easy estimation from time series data. For these reasons, we focus on the methods in [10] to empirically estimate Φ from an observed time series. The building block of integrated information is effective information, ϕ. Effective information quantifies how much better a system X is at predicting its own future (or decoding its own past) after a time τ when it is considered where K is a normalisation factor to avoid biasing Φ to excessively unbalanced bipartitions. Defined this way, Φ can be understood as the minimum information loss incurred by splitting the system into two subsystems. It quantifies the collective emergent behaviour that is present in the whole system but not in any bipartition. III. RESULTS We ran 1500 simulations with values of β distributed uniformly at random in the range [0, 2π) using RK4 with a stepsize of 0.05 for numerical integration. Each simula- tion was run for 5 × 106 timesteps, of which the first 104 are discarded to avoid effects from transient states. All information-theoretic measures are reported in bits. We first study the system from a purely dynamical per- spective, following the analysis in [4]. Global synchrony and metastability are shown in Fig. 1. The first character- istic we observe is that there are two well differentiated dynamical regimes – one of hypersynchronisation and one of complete desynchronisation, with strong metastability appearing in the narrow transition bands between one and the other. It is in this transition region where the oscillators oper- ate in a critical regime poised between order and disorder and complex phenomena appear. As the system moves from desynchronisation to full synchronisation there is a sharp increase in metastability, followed by a smoother decrease as the system becomes hypersynchronised. In the region 0 < β < π/8, the system remains in a complex equilibrium between an ordered and a disordered phase. 3 FIG. 1. Global synchrony and metastability for different phase lags β for the whole [0, 2π) range (bottom) and around the critical transition region (top). Rapid increase of metastability marks the onset of the phase transition. Note the different β ranges in both plots. FIG. 2. Integrated information Φ and coalition entropy Hc in the phase transition. Within the broad region between order and disorder in which Hc rises there is a narrower band in which complex spatiotemporal patterns generate high Φ. A. Information-theoretic analysis One feature of Φ (and of any other information-theoretic measure), compared to λ, is that it does not need to be cal- culated directly from the state of the system. According to its definition, Φ is substrate-agnostic, meaning that the relevant quantity for the calculation of Φ is not the physi- cal state of the system, but some informational state – the configuration of the system that we consider to contain information. Therefore, we must define an informational state mapping, that extracts the information-bearing sym- bols from the physical state of the system. Although calculating Φ on the real-valued phases is possible, for simplicity we choose the coalition configura- tion of the system as the informational state, defined as the set of communities that are highly internally synchro- nised. To calculate the coalition configuration at time t we calculate Rc(t) of each community and threshold it, such that (cid:40) X c t = if Rc(t) > γ 1 0 otherwise. We refer to γ as the coalition threshold. After calculat- ing the coalitions, the history of the system is reduced to a time series with 8 binary variables. Having a multivariate discrete time series, it is now tractable to compute Φ. By default, we use γ = 0.8 in all our analyses shown here. As depicted in Fig. 2, Φ shows a similar behaviour to λ – it peaks in the transition regions and shrinks in the fully ordered and the fully disordered regimes. We also compare Φ with arguably the simplest information- theoretic measure – entropy H. The entropy of the state of the network calculated on the coalitions Xt forms the coalition entropy Hc. Both Φ and Hc peak precisely at the same point. Al- though both measures depend on the chosen coalition threshold γ, the results are qualitatively the same for a wide range of thresholds. Although it peaks in the same region as λ and Hc, we note that Φ reveals new properties of the system by virtue of incorporating temporal information in its definition. That is, to have a high Φ a system must exhibit complex spatial and temporal patterns. We can verify this by performing a random time-shuffle on the time series. This shuffling leaves λ and Hc unaltered, as they don't explicitly depend on time correlations, but has a high impact on Φ, which shrinks to zero. This indicates that Φ is sensitive to properties of the system that are not reflected by other measures. Furthermore, Φ can be used to investigate the behaviour of the system at multiple timescales. Figure 3 shows the behaviour of Φ for several values of τ , and compares it with standard time-delayed mutual information (TDMI) I(Xt−τ , Xt). The first thing we note is that Φ and TDMI have oppo- site trends with τ . TDMI decreases for longer timescales while Φ increases. At short timescales the system is highly predictable – thus the high TDMI – but this short-term evolution does not involve any system-wide interaction – thus the low Φ. Furthermore, at these timescales Φ is negative, which can be interpreted as an indication of redundancy [17] in the evolution of the system: the parts share some information, such that they separately con- tain more information about their past than the whole system about its past. For larger τ TDMI decreases, as the evolution of the system becomes harder to track using the coalition configuration. Simultaneously, Φ becomes higher, indicating that the remaining TDMI has a stronger integrated component that is not accounted for by the TDMI of the partitions of the system. Overall, we see a 00.20.40.600.020.04Metastabilityλ00.20.40.60.81Synchronyξλξ0π2π3π200.020.04Phaselagβλ00.20.40.60.81ξ00.10.20.30.40.500.511.522.5PhaselagβIntegratedInformationΦ02468CoalitionEntropyHcΦHc 4 Φ. This means that Φ is sensitive to more complicated dynamic patterns than the other measures considered, and is in that sense more discriminating. We note that λ is a community-local quantity – that is, the calculation of λc for each community is independent of the rest. Conversely, Φ relies exclusively on the irreducible interaction between communities. These two quantities are nevertheless intrinsically related, insofar as internal variability enables the system to visit a larger repertoire of states in which system-wide interaction can take place. B. Robustness of Φ against measurement noise We will now consider the impact of measurement noise on Φ, wherein the system runs unchanged but our record- ing of it is imperfect. For this experiment we run the (deterministic) simulation as presented in the previous section and take the binary time series of coalition con- figurations. We then emulate the effect of uncorrelated measurement noise by flipping each bit in the time series with probability p, yielding the corrupted time series X. Finally we recalculate Φ on the corrupted time series, and show the results in Fig. 4. To quantify how fast Φ changes we calculate the ratio between the corrupted and the original time series, η = Φ[ X, τ ] Φ[X, τ ] . (7) In order to avoid instabilities as Φ[X, τ ] gets close to zero, we calculate η only in the region within 0.5 rad of the centre of the peak, where Φ[X, τ ] is large. The inset of Fig. 4 shows the mean and standard deviation of η at different noise levels p. FIG. 4. Integrated information Φ for different levels of mea- surement noise p. Inset: (blue) Mean and variance of the ratio η between Φ of the corrupted and the original time series. (red) Exponential fit η = exp(−p/(cid:96)), with (cid:96) ≈ 0.04. FIG. 3. Integrated information Φ and time-delayed mutual information I(Xt−τ , Xt) for several timescales τ . See text for details. clear trend of TDMI diminishing at longer timescales but becoming progressively more integrated in nature. It might seem counterintuitive to the reader that both TDMI and Φ converge to a non-zero value for arbitrarily high τ . However, recall that the system is deterministic, so there is no reason to expect TDMI to vanish even in the τ → ∞ limit. Perfect knowledge of all of the oscillators' phases at any given time step is enough to reconstruct the whole history of the system. The reason why we see a decreasing TDMI is because the informa- tional state we chose (the coalition configuration) is not descriptive enough to capture the evolution of the system perfectly. That is, this is not a result of stochasticity, but of degeneracy. Put another way, the TDMI of the whole system re- mains non-zero in the τ → ∞ limit because we are dealing with a causally closed system. Again, if we considered the totality of the system's state (i.e. the phases θi) then TDMI would be maximal and constant for any τ . In contrast, when considering the TDMI of any partition M we start dealing with a causally open system, since one partition is affected by the other. This effectively intro- duces stochasticity in our observations Mt, which does cause TDMI of the partition to vanish when τ → ∞. This explains that the TDMI of the whole system converges to a non-zero value for large τ while the TDMI of any partition fades to zero, leaving a positive Φ. Finally, it is interesting to combine the insights from the dynamical and information-theoretic analyses. Inspecting Figs. 1 and 2 we see that the peak in Φ is much narrower than the peaks in λ and Hc. While some values of β do give rise to non-trivial dynamics, it is only at the centre of the critical region that these dynamics give rise to integration. A certain degree of internal variability is necessary to establish integrated information, but not all configurations with high internal variability lead to a high 00.10.20.30.40.50.01.02.0Φ[X,τ]τ=1τ=10τ=10000.10.20.30.40.50.02.04.06.0PhaselagβI(Xt−τ,Xt)00.10.20.30.40.500.511.52PhaselagβΦ[X,τ]p=0.00p=0.02p=0.04p=0.0600.050.10.150.200.51pη 5 We find that Φ monotonically decays with p, reflecting the gradual loss of the precise spatiotemporal patterns characteristic of the system. The distortion has a greater effect on time series with greater Φ, but preserves the dominant peak in β ≈ 0.15. The inset shows that both the mean and variance of η decay as a clean exponential with p. Φ is highly sensitive to noise and undergoes a rapid decline, as a measurement noise of 5% wipes out 70% of the perceived integrated information of the system. IV. CONCLUSION We have presented a community-structured network of Kuramoto oscillators and discussed their collective behaviour in terms of metastability [4] and integrated information [9]. We showed that the system undergoes a phase transition whose critical region presents a sharp, clear peak of integrated information Φ that coincides with a strong increase in metastability. To our knowledge, this is the first description of a dynamical system in which the three major complexity indicators of criticality, high metastability, and high integrated information all appear. The resulting confluence of two major research directions in complexity science suggests that this is a system that merits further study. In the context of the present model, the high internal variability of the system's components enables system- wide interaction, which in turn leads to high Φ. As we have seen, the system presents a region of high metasta- bility, but notably it is only within an even narrower band that we find strong integrated information. Moreover, as we have also seen, shuffling the time series data preserves the peak of metastability, despite the fact that the result is meaningless. By contrast, Φ only peaks when the relevant temporal structure is present in the data. In this way we provide evidence that complex dynamics – as quantified by the metastability index λ – are a necessary but not sufficient condition for complex information processing – as quantified by integrated information Φ. Dynamical and information-theoretic measures provide different lenses through which we can understand a sys- tem, and offer complementary views of its behaviour. Our findings support the claim that Φ, despite having some theoretical drawbacks [11], is a valuable tool for understanding complex spatial and temporal behaviour in dynamical systems, particularly when combined with other analysis techniques. [1] A. Pikovsky, M. Rosenblum, and J. Kurths, Synchro- nization: A Universal Concept in Nonlinear Sciences (Cambridge University Press, Cambridge, 2001) p. 432. [2] Y. Kuramoto, Chemical Oscillations, Waves and Turbu- lence (Dover Publications, 1984) p. 164. [3] M. J. Panaggio and D. M. Abrams, Nonlinearity 28, R67 (2015), arXiv:1403.6204. [4] M. Shanahan, Chaos 20, 013108 (2010), arXiv:0908.3881. [5] J. Cabral, E. Hugues, O. Sporns, and G. Deco, NeuroIm- age 57, 130 (2011). [6] P. J. Hellyer, G. Scott, M. Shanahan, D. J. Sharp, and R. Leech, The Journal of Neuroscience 35, 9050 (2015). [7] M. Schartner, A. Seth, Q. Noirhomme, M. Boly, M.-A. Bruno, S. Laureys, and A. B. Barrett, PloS One 10, e0133532 (2015). [8] M. Shanahan, Physical Review E 78, 041924 (2008). [9] D. Balduzzi and G. Tononi, PLoS Computational Biology 4, e1000091 (2008). [10] A. B. Barrett and A. K. Seth, PLoS Computational Biol- ogy 7, e1001052 (2011). [11] V. Griffith, arXiv:1401.0978. [12] M. Tegmark, arXiv:1601.02626. [13] D. M. Abrams, R. Mirollo, S. H. Strogatz, and D. A. Wiley, Physical Review Letters 101, 084103 (2008). [14] J.-R. King, J. D. Sitt, F. Faugeras, B. Rohaut, I. El Karoui, L. Cohen, L. Naccache, and S. Dehaene, Current Biology 23, 1914 (2013). [15] M. Oizumi, L. Albantakis, and G. Tononi, PLoS Compu- tational Biology 10, e1003588 (2014). [16] G. Tononi, Archives Italiennes de Biologie 150, 56 (2012). [17] A. B. Barrett, arXiv:1411.2832.
1608.08040
2
1608
2017-03-08T18:14:34
Rules and mechanisms for efficient two-stage learning in neural circuits
[ "q-bio.NC", "physics.bio-ph" ]
Trial-and-error learning requires evaluating variable actions and reinforcing successful variants. In songbirds, vocal exploration is induced by LMAN, the output of a basal ganglia-circuit that also contributes a corrective bias to the vocal output. This bias is gradually consolidated in RA, a motor cortex analogue downstream of LMAN. We develop a new model of such two-stage learning. Using stochastic gradient descent, we derive how the activity in 'tutor' circuits (e.g., LMAN) should match plasticity mechanisms in 'student' circuits (e.g., RA) to achieve efficient learning. We further describe a reinforcement learning framework through which the tutor can build its teaching signal. We show that mismatches between the tutor signal and the plasticity mechanism can impair learning. Applied to birdsong, our results predict the temporal structure of the corrective bias from LMAN given a plasticity rule in RA. Our framework can be applied predictively to other paired brain areas showing two-stage learning.
q-bio.NC
q-bio
Rules and mechanisms for efficient two-stage learning in neural circuits Tiberiu Te¸sileanu1,2, Bence Olveczky3, and Vijay Balasubramanian1,2 1Initiative for the Theoretical Sciences, The Graduate Center, CUNY, New York, NY 10016 2David Rittenhouse Laboratories, University of Pennsylvania, Philadelphia, PA 19104 3Department of Organismic and Evolutionary Biology and Center for Brain Science, Harvard University, Cambridge, MA 02138 November 9, 2018 Abstract Trial-and-error learning requires evaluating variable actions and reinforcing successful variants. In songbirds, vocal exploration is induced by LMAN, the output of a basal ganglia-circuit that also contributes a corrective bias to the vocal output. This bias is gradually consolidated in RA, a motor cortex analogue downstream of LMAN. We develop a new model of such two-stage learning. Using stochastic gradient descent, we derive how the activity in 'tutor' circuits (e.g., LMAN) should match plasticity mechanisms in 'student' circuits (e.g., RA) to achieve efficient learning. We further describe a reinforcement learning framework through which the tutor can build its teaching signal. We show that mismatches between the tutor signal and the plasticity mechanism can impair learning. Applied to birdsong, our results predict the temporal structure of the corrective bias from LMAN given a plasticity rule in RA. Our framework can be applied predictively to other paired brain areas showing two-stage learning. 1 Introduction Two-stage learning has been described in a variety of different contexts and neural circuits. During hippocampal memory consolidation, recent memories, that are dependent on the hippocampus, are transferred to the neocortex for long-term storage (Frankland and Bontempi 2005). Similarly, the rat motor cortex provides essential input to sub-cortical circuits during skill learning, but then becomes dispensable for executing certain skills (Kawai et al. 2015). A paradigmatic example of two-stage learning occurs in songbirds learning their courtship songs (Andalman and Fee 2009; Turner and Desmurget 2010; Warren et al. 2011). Zebra finches, commonly used in birdsong research, learn their song from their fathers as juveniles and keep the same song for life (Immelmann 1969). The birdsong circuit has been extensively studied; see Figure 1A for an outline. Area HVC is a timebase circuit, with projection neurons that fire sparse spike bursts in precise synchrony with the song (Hahnloser, Kozhevnikov, and Fee 2002; Lynch et al. 2016; Picardo et al. 2016). A population of neurons from HVC projects to the robust nucleus of the arcopallium (RA), a pre-motor area, which then projects to motor neurons controlling respiratory and syringeal muscles (Leonardo and Fee 2005; Simpson and Vicario 1990; Yu and Margoliash 1996). A second input to RA comes from the lateral magnocellular nucleus of the anterior nidopallium (LMAN). Unlike HVC and RA activity patterns, LMAN spiking is highly variable across different renditions of the song (Kao, Wright, and Doupe 2008; Olveczky, Andalman, and Fee 2005). LMAN is the output of the anterior forebrain pathway, a circuit involving the song-specialized basal ganglia (Perkel 2004). Because of the variability in its activity patterns, it was thought that LMAN's role was simply to inject variability into the song ( Olveczky, Andalman, and Fee 2005). The resulting vocal experimentation would enable reinforcement-based learning. For this reason, prior models tended to treat LMAN as a pure Poisson noise generator, and assume that a reward signal is received directly in RA (Fiete, Fee, and Seung 2007). More recent evidence, however, suggests that the reward signal reaches Area X, the song-specialized basal ganglia, 1 rather than RA (Gadagkar et al. 2016; Hoffmann et al. 2016; Kubikova and Kost´al 2010). Taken together with the fact that LMAN firing patterns are not uniformly random, but rather contain a corrective bias guiding plasticity in RA (Andalman and Fee 2009; Warren et al. 2011), this suggests that we should rethink our models of song acquisition. Here we build a general model of two-stage learning where one neural circuit "tutors" another. We develop a formalism for determining how the teaching signal should be adapted to a specific plasticity rule, to best instruct a student circuit to improve its performance at each learning step. We develop analytical results in a rate based model, and show through simulations that the general findings carry over to realistic spiking neurons. Applied to the vocal control circuit of songbirds, our model reproduces the observed changes in the spiking statistics of RA neurons as juvenile birds learn their song. Our framework also predicts how the LMAN signal should be adapted to properties of RA synapses. This prediction can be tested in future experiments. Our approach separates the mechanistic question of how learning is implemented from what the resulting learning rules are. We nevertheless demonstrate that a simple reinforcement learning algorithm suffices to implement the learning rule we propose. Our framework makes general predictions for how instructive signals are matched to plasticity rules whenever information is transferred between different brain regions. Figure 1: Relation between the song system in zebra finches and our model. A. Diagram of the major brain regions involved in birdsong. B. Conceptual model inspired by the birdsong system. The line from output to tutor is dashed because the reinforcement signal can reach the tutor either directly or, as in songbirds, indirectly. C. Plasticity rule measured in bird RA (measurement done in slice). When an HVC burst leads an LMAN burst by about 100 ms, the HVC -- RA synapse is strengthened, while coincident firing leads to suppression. Figure adapted from (Mehaffey and Doupe 2015). D. Plasticity rule in our model that mimics the Mehaffey and Doupe (2015) rule. 2 motorHVCRALMANArea XDLMconductorstudenttutoroutputreinforcementABCD 2 Results 2.1 Model We considered a model for information transfer that is composed of three sub-circuits: a conductor, a student, and a tutor (see Figure 1B). The conductor provides input to the student in the form of temporally precise patterns. The goal of learning is for the student to convert this input to a predefined output pattern. The tutor provides a signal that guides plasticity at the conductor -- student synapses. For simplicity, we assumed that the conductor always presents the input patterns in the same order, and without repetitions. This allowed us to use the time t to label input patterns, making it easier to analyze the on-line learning rules that we studied. This model of learning is based on the logic implemented by the vocal circuits of the songbird (Figure 1A). Relating this to the songbird, the conductor is HVC, the student is RA, and the tutor is LMAN. The song can be viewed as a mapping between clock-like HVC activity patterns and muscle-related RA outputs. The goal of learning is to find a mapping that reproduces the tutor song. Birdsong provides interesting insights into the role of variability in tutor signals. If we focus solely on information transfer, the tutor output need not be variable; it can deterministically provide the best instructive signal to guide the student. This, however, would require the tutor to have a detailed model of the student. More realistically, the tutor might only have access to a scalar representation of how successful the student rendition of the desired output is, perhaps in the form of a reward signal. A tutor in this case has to solve the so-called 'credit assignment problem' -- it needs to identify which student neurons are responsible for the reward. A standard way to achieve this is to inject variability in the student output and reinforcing the firing of neurons that precede reward (see for example (Fiete, Fee, and Seung 2007) in the birdsong context). Thus, in our model, the tutor has a dual role of providing both an instructive signal and variability, as in birdsong. We described the output of our model using a vector ya(t) where a indexed the various output channels (Figure 2A). In the context of motor control a might index the muscle to be controlled, or, more abstractly, different features of the motor output, such as pitch and amplitude in the case of birdsong. The output ya(t) was a function of the activity of the student neurons sj(t). The student neurons were in turn driven by the activity of the conductor neurons ci(t). The student also received tutor signals to guide plasticity; in the songbird, the guiding signals for each RA neuron come from several LMAN neurons (Canady et al. 1988; Garst-Orozco, Babadi, and Olveczky 2014; Herrmann and Arnold 1991). In our model, we summarized the net input from the tutor to the jth student neuron as a single function gj(t). We started with a rate-based implementation of the model (Figure 2A) that was analytically tractable but averaged over tutor variability. We further took the neurons to be in a linear operating regime (Figure 2A) away from the threshold and saturation present in real neurons. We then relaxed these conditions and tested our results in spiking networks with initial parameters selected to imitate measured firing patterns in juvenile birds prior to song learning. The student circuit in both the rate-based and spiking models included a global inhibitory signal that helped to suppress excess activity driven by ongoing conductor and tutor input. Such recurrent inhibition is present in area RA of the bird (Spiro, Dalva, and Mooney 1999). In the spiking model we implemented the suppression as an activity-dependent inhibition, while for the analytic calculations we used a constant negative bias for the student neurons. 2.2 Learning in a rate-based model Learning in our model was enabled by plasticity at the conductor -- student synapses that was modulated by signals from tutor neurons (Figure 2B). Many different forms of such hetero-synaptic plasticity have been observed. For example, in rate-based synaptic plasticity high tutor firing rates lead to synaptic potentiation and low tutor firing rates lead to depression (Chistiakova, Bannon, et al. 2014; Chistiakova and Volgushev 2009). In timing-dependent rules, such as the one recently measured by Mehaffey and Doupe (2015) in slices of zebra finch RA (see Figure 1C), the relative arrival times of spike bursts from different input pathways set the sign of synaptic change. To model learning that lies between these rate and timing-based extremes, we introduced a class of plasticity rules governed by two parameters α and β (see also Methods and Figure 2B): (cid:21) , )/τ1 − β τ2 e −(t−t(cid:48) )/τ2 (1) dWij dt ci(t) = 0 = ηci(t)(cid:0)gj(t) − θ(cid:1) , (cid:20) α (cid:90) t (cid:48) (cid:48) dt ci(t ) −(t−t(cid:48) e τ1 3 Figure 2: Schematic representation of our rate-based model. A. Conductor neurons fire precisely-timed bursts, similar to HVC neurons in songbirds. Conductor and tutor activities, c(t) and g(t), provide excitation to student neurons, which integrate these inputs and respond linearly, with activity s(t). Student neurons also receive a constant inhibitory input, xinh. The output neurons linearly combine the activities from groups of student neurons using weights Maj. The linearity assumptions were made for mathematical convenience but are not essential for our qualitative results (see Appendix). B. The conductor -- student synaptic weights Wij are updated based on a plasticity rule that depends on two parameters, α and β, and two timescales, τ1 and τ2 (see eq. (1) and Methods). The tutor signal enters this rule as a deviation from a constant threshold θ. The figure shows how synaptic weights change (∆W ) for a student neuron that receives a tutor burst and a conductor burst separated by a short lag. Two different choices of plasticity parameters are illustrated in the case when the threshold θ = 0. C. The amount of mismatch between the system's output and the target output is quantified using a loss (error) function. The figure sketches the loss landscape obtained by varying the synaptic weights Wij and calculating the loss function in each case (only two of the weight axes are shown). The blue dot shows the lowest value of the loss function, corresponding to the best match between the motor output and the target, while the orange dot shows the starting point. The dashed line shows how learning would proceed in a gradient descent approach, where the weights change in the direction of steepest descent in the loss landscape. 4 conductorstudent∑tutoroutput-400400-400400tutorconductorerrorABClosslandscapesynapticplasticity(loss function)convolution∑∑∑∑∑t (ms)t (ms)(ms)(ms) where Wij is the weight of the synapse from the ith conductor to the jth student neuron, η is a learning rate, θ is a threshold on the firing rate of tutor neurons, and τ1 and τ2 are timescales associated with the plasticity. This is similar to an STDP rule, except that the dependence on postsynaptic activity was replaced by dependence on the input from the tutor. Thus plasticity acts heterosynaptically, with activation of the tutor -- student synapse controlling the change in the conductor -- student synaptic weight. The timescales τ1 and τ2, as well as the coefficients α and β, can be thought of as effective parameters describing the plasticity observed in student neurons. As such, they do not necessarily have a simple correspondence in terms of the biochemistry of the plasticity mechanism, and the framework we describe here is not specifically tied to such an interpretation. If we set α or β to zero in our rule, eq. (1), the sign of the synaptic change is determined solely by the firing rate of the tutor gj(t) as compared to a threshold, reproducing the rate rules observed in experiments. When α/β ≈ 1, if the conductor leads the tutor, potentiation occurs, while coincident signals lead to depression (Figure 2B), which mimics the empirical findings from (Mehaffey and Doupe 2015). For general α and β, the sign of plasticity is controlled by both the firing rate of the tutor relative to the baseline, and by the relative timing of tutor and conductor. The overall scale of the parameters α and β can be absorbed into the learning rate η and so we set α − β = 1 in all our simulations without loss of generality (see Methods). Note that if α and β are both large, it can be that α − β = 1 and α/β ≈ 1 also, as needed to realize the Mehaffey and Doupe (2015) curve. We can ask how the conductor -- student weights Wij (Figure 2A) should change in order to best improve the output ya(t). We first need a loss function L that quantifies the distance between the current output ya(t) and the target ¯ya(t) (Figure 2C). We used a quadratic loss function, but other choices can also be incorporated into our framework (see Appendix). Learning should change the synaptic weights so that the loss function is minimized, leading to a good rendition of the targeted output. This can be achieved by changing the synaptic weights in the direction of steepest descent of the loss function (Figure 2C). We used the synaptic plasticity rule from eq. (1) to calculate the overall change of the weights, ∆Wij, over the course of the motor program. This is a function of the time course of the tutor signal, gj(t). Not every choice for the tutor signal leads to motor output changes that best improve the match to the target. Imposing the condition that these changes follow the gradient descent procedure described above, we derived the tutor signal that was best matched to the student plasticity rule (detailed derivation in Methods). The result is that the best tutor for driving gradient descent learning must keep track of the motor error (cid:88) j(t) = Maj(ya(t) − ¯ya(t)) (2) (3) integrated over the recent past gj(t) = θ − a ζ 1 (cid:90) t α − β τtutor 0 (cid:48) j(t )e −(t−t(cid:48) (cid:48) )/τtutor dt , where Maj are the weights describing the linear relationship between student activities and motor outputs (Figure 2A) and ζ is a learning rate. Moreover, for effective learning, the timescale τtutor appearing in eq. (3), which quantifies the timescale on which error information is integrated into the tutor signal, should be related to the synaptic plasticity parameters according to τtutor = τ ∗ tutor ≡ is the optimal timescale for the error integration. τ ∗ tutor , where ατ1 − βτ2 α − β (4) In short, motor learning with a heterosynaptic plasticity rule requires convolving the motor error with a kernel whose timescale is related to the structure of the plasticity rule, but is otherwise independent of the motor program.1 As explained in more detail in Methods, this result is derived in an approximation that assumes that the tutor signal does not vary significantly over timescales of the order of the student timescales τ1 and τ2. Given eq. (4), this implies that we are assuming τtutor (cid:29) τ1,2. This is a reasonable approximation because variations in the tutor signal that are much faster than the student timescales τ1,2 have little effect on learning since the plasticity rule (1) blurs conductor inputs over these timescales. 1We thank the referees for suggesting this way of describing our results. 5 2.3 Matched vs. unmatched learning Our rate-based model predicts that when the timescale on which error information is integrated into the tutor signal (τtutor) is matched to the student plasticity rule as described above, learning will proceed efficiently. A mismatched tutor should slow or disrupt convergence to the desired output. To test this, we numerically simulated the birdsong circuit using the linear model from Figure 2A with a motor output ya filtered to more realistically reflect muscle response times (see Methods). We selected plasticity rules as described in eq. (1) and Figure 2B and picked a target output pattern to learn. The target was chosen to resemble recordings of air-sac pressure from singing zebra finches in terms of smoothness and characteristic timescales (Veit, Aronov, and Fee 2011), but was otherwise arbitrary. In our simulations, the output typically involved two different channels, each with its own target, but for brevity, in figures we typically showed the output from only one of these. For our analytical calculations, we made a series of assumptions and approximations meant to enhance tractability, such as linearity of the model and a focus on the regime τtutor (cid:29) τ1,2. These constraints can be lifted in our simulations, and indeed below we test our numerical model in regimes that go beyond the approximations made in our derivation. In many cases, we found that the basic findings regarding tutor -- student matching from our analytical model remain true even when some of the assumptions we used to derive it no longer hold. We tested tutors that were matched or mismatched to the plasticity rule to see how effectively they instructed the student. Figure 3A and online Video 1 show convergence with a matched tutor when the sign of plasticity is determined by the tutor's firing rate. We see that the student output rapidly converged to the target. Figure 3B and online Video 2 show convergence with a matched tutor when the sign of plasticity is largely determined by the relative timing of the tutor signal and the student output. We see again that the student converged steadily to the desired output, but at a somewhat slower rate than in Figure 3A. To test the effects of mismatch between tutor and student, we used tutors with timescales that did not match eq. (4). All student plasticity rules had the same effective time constants τ1 and τ2, but different parameters α and β (see eq. (1)), subject to the constraint α − β = 1 described in section 2.2. Different tutors had different memory time scales τtutor (eq. (3)). Figures 3C and 3D demonstrate that learning was more rapid for well-matched tutor-student pairs (the diagonal neighborhood, where τtutor ≈ τ∗ tutor). When the tutor error integration timescale was shorter than the matched value in eq. (4), τtutor < τ∗ tutor, learning was often completely disrupted (many pairs below the diagonal in Figures 3C and 3D). When the tutor error integration timescale was longer than the matched value in eq. (4), τtutor > τ∗ tutor learning was slowed down. Figure 3C also shows that a certain amount of mismatch between the tutor error integration timescale τtutor and the matched timescale τ∗ tutor implied by the student plasticity rule is tolerated by the system. Interestingly, the diagonal band over which learning is effective in Figure 3C is roughly of constant width -- note that the scale on both axes is logarithmic, so that this means that the tutor error integration timescale τtutor has to be within a constant factor of the optimal timescale τ∗ tutor for good learning. We also see that the breakdown in learning is more abrupt when τtutor < τ∗ tutor than in the opposite regime. An interesting feature of the results from Figures 3C, 3D is that the difference in performance between matched and mismatched pairs becomes less pronounced for timescales shorter than about 100 ms. This is due to the fact that the plasticity rule (eq. (1)) implicitly smooths over timescales of the order of τ1,2, which in our simulations were equal to τ1 = 80 ms, τ2 = 40 ms. Thus, variations of the tutor signal on shorter timescales have little effect on learning. Using different values for the effective timescales τ1,2 describing the plasticity rule can increase or decrease the range of parameters over which learning is robust against tutor -- student mismatches (see Appendix). 2.4 Robust learning with nonlinearities In the model above, firing rates for the tutor were allowed to grow as large as necessary to implement the most efficient learning. However, the firing rates of realistic neurons typically saturate at some fixed bound. To test the effects of this nonlinearity in the tutor, we passed the ideal tutor activity (3) through a sigmoidal nonlinearity, (cid:90) t 6 gj(t) = θ − ρ tanh α − β τtutor 0 ζ 1 (cid:48) j(t −(t−t(cid:48) )e (cid:48) )/τtutor dt . (5) where 2ρ is the range of firing rates. We typically chose θ = ρ = 80 Hz to constrain the rates to the range 0 -- 160 Hz (Garst-Orozco, Babadi, and Olveczky 2014; Olveczky, Andalman, and Fee 2005). Learning slowed Figure 3: Learning with matched or mismatched tutors in rate-based simulations. A. Error trace showing how the average motor error evolved with the number of repetitions of the motor program for a rate-based (α = 0) plasticity rule paired with a matching tutor. (See online Video 1.) B. The error trace and final motor output shown for a timing-based learning rule matched by a tutor with a long integration timescale. (See online Video 2.) In both A and B the inset shows the final motor output for one of the two output channels (thick orange line) compared to the target output for that channel (dotted black line). The output on the first rendition and at two other stages of learning indicated by orange arrows on the error trace are also shown as thin orange lines. C. Effects of mismatch between student and tutor on reproduction accuracy. The heatmap shows the final reproduction error of the motor output after 1000 learning cycles in a rate-based simulation where a student with parameters α, β, τ1, and τ2 was paired with a tutor with memory timescale τtutor. On the y axis, τ1 and τ2 were kept fixed at 80 ms and 40 ms, respectively, while α and β were varied (subject to the constraint α − β = 1; see text). Different choices of α and β lead to different optimal timescales τ∗ tutor according to eq. (4). The diagonal elements correspond to matched tutor and student, τtutor = τ∗ tutor. Note that the color scale is logarithmic. D. Error evolution curves as a function of the mismatch between student and tutor. Each plot shows how the error in the motor program changed during 1000 learning cycles for the same conditions as those shown in the heatmap. The region shaded in light pink shows simulations where the mismatch between student and tutor led to a deteriorating instead of improving performance during learning. Video 1: Evolution of motor output during learning in a rate-based simulation using a rate-based (α = 0) plasticity rule paired with a matching tutor. This video relates to Figure 3A. Video 2: Evolution of motor output during learning in a rate-based simulation using a timing-based (α ≈ β) plasticity rule paired with a matching tutor. This video relates to Figure 3B. 7 102040801603206401280256051201024020480τtutor102040801603206401280256051201024020480τ∗tutor=ατ1−βτ2α−β102040801603206401280256051201024020480τtutor102040801603206401280256051201024020480τ∗tutor=ατ1−βτ2α−β0.501.002.005.0010.00050100150200250repetitions02468101214errorα=7.0,β=6.0,τtutor=320.0time010203040506070outputtargetoutput050100150200250repetitions02468101214errorα=0.0,β=−1.0,τtutor=40.0time01020304050607080outputtargetoutput0600ABCD0600worsebetter down with this change (Figure 4A and online Video 3) as a result of the tutor firing rates saturating when the mismatch between the motor output and the target output was large. However, the accuracy of the final rendition was not affected by saturation in the tutor (Figure 4A, inset). An interesting effect occurred when the firing rate constraint was imposed on a matched tutor with a long memory timescale. When this happened and the motor error was large, the tutor signal saturated and stopped growing in relation to the motor error before the end of the motor program. In the extreme case of very long integration timescales, learning became sequential: early features in the output were learned first, before later features were addressed, as in Figure 4B and online Video 4. This is reminiscent of the learning rule described in (Memmesheimer et al. 2014). Figure 4: Effects of adding a constraint on the tutor firing rate to the simulations. A. Learning was slowed down by the firing rate constraint, but the accuracy of the final rendition stayed the same (inset, shown here for one of two simulated output channels). Here α = 0, β = −1, and τtutor = τ∗ tutor = 40 ms. (See online Video 3.) B. Sequential learning occurred when the firing rate constraint was imposed on a matched tutor with a long memory scale. The plots show the evolution of the motor output for one of the two channels that were used in the simulation. Here α = 24, β = 23, and τtutor = τ∗ tutor = 1000 ms. (See online Video 4.) Video 3: Effects of adding a constraint on tutor firing rates on the evolution of motor output during learning in a rate-based simulation. The plasticity rule here was rate-based (α = 0). This video relates to Figure 4A. Video 4: Evolution of the motor output showing sequential learning in a rate-based simulation, which occurs when the firing rate constraint is imposed on a tutor with a long memory timescale. This video relates to Figure 4B. Nonlinearities can similarly affect the activities of student neurons. Our model can be readily extended to describe efficient learning even in this case. The key result is that for efficient learning to occur, the synaptic plasticity rule should depend not just on the tutor and conductor, but also on the activity of the postsynaptic student neurons (details in Appendix). Such dependence on postsynaptic activity is commonly seen in experiments (Chistiakova, Bannon, et al. 2014; Chistiakova and Volgushev 2009). The relation between student neuron activations sj(t) and motor outputs ya(t) (Figure 2A) is in general also nonlinear. Compared to the linear assumption that we used, the effect of a monotonic nonlinearity, 8 0600time020406080Stage1targetoutput0600time020406080Stage17targetoutput0600time020406080Stage34targetoutput050100150200250repetition05101520errortimeoutputtargetnoconstraintwithconstraintAB0600 ya = Na((cid:80) j Majsj), with Na an increasing function, is similar to modifying the loss function L, and does not significantly change our results (see Appendix). We also checked that imposing a rectification constraint that conductor -- student weights Wij must be positive does not modify our results either (see Appendix). This shows that our model continues to work with biologically realistic synapses that cannot change sign from excitatory to inhibitory during learning. 2.5 Spiking neurons and birdsong To apply our model to vocal learning in birds, we extended our analysis to networks of spiking neurons. Juvenile songbirds produce a "babble" that converges through learning to an adult song strongly resembling the tutor song. This is reflected in the song-aligned spiking patterns in pre-motor area RA, which become more stereotyped and cluster in shorter, better-defined bursts as the bird matures (Figure 5A). We tested whether our model could reproduce key statistics of spiking in RA over the course of song learning. In this context, our theory of efficient learning, derived in a rate-based scenario, predicts a specific relation between the teaching signal embedded in LMAN firing patterns, and the plasticity rule implemented in RA. We tested whether these predictions continued to hold in the spiking context. Following the experiments of Hahnloser, Kozhevnikov, and Fee (2002), we modeled each neuron in HVC (the conductor) as firing one short, precisely timed burst of 5-6 spikes at a single moment in the motor program. Thus the population of HVC neurons produced a precise timebase for the song. LMAN (tutor) neurons are known to have highly variable firing patterns that facilitate experimentation, but also contain a corrective bias (Andalman and Fee 2009). Thus we modeled LMAN as producing inhomogeneous Poisson spike trains with a time-dependent firing rate given by eq. (5) in our model. Although biologically there are several LMAN neurons projecting to each RA neuron, we again simplified by "summing" the LMAN inputs into a single, effective tutor neuron, similarly to the approach in (Fiete, Fee, and Seung 2007). The LMAN-RA synapses were modeled in a current-based approach as a mixture of AMPA and NMDA receptors, following the songbird data (Garst-Orozco, Babadi, and Olveczky 2014; Stark and Perkel 1999). The initial weights for all synapses were tuned to produce RA firing patterns resembling juvenile birds ( Olveczky, Otchy, et al. 2011), subject to constraints from direct measurements in slice recordings (Garst-Orozco, Babadi, and Olveczky 2014) (see Methods for details, and Figure 5B for a comparison between neural recordings and spiking in our model). In contrast to the constant inhibitory bias that we used in our rate-based simulations, for the spiking simulations we chose an activity-dependent global inhibition for RA neurons. We also tested that a constant bias produced similar results (see Appendix). Synaptic strength updates followed the same two-timescale dynamics that was used in the rate-based models (Figure 2B). The firing rates ci(t) and gj(t) that appear in the plasticity equation were calculated in the spiking model by filtering the spike trains from conductor and tutor neurons with exponential kernels. The synaptic weights were constrained to be non-negative. (See Methods for details.) As long as the tutor error integration timescale was not too large, learning proceeded effectively when the tutor error integration timescale and the student plasticity rule were matched (see Figure 5C and online Video 5), with mismatches slowing down or abolishing learning, just as in our rate-based study (compare Figure 5D with Figure 3C). The rate of learning and the accuracy of the trained state were lower in the spiking model compared to the rate-based model. The lower accuracy arises because the tutor neurons fire stochastically, unlike the deterministic neurons used in the rate-based simulations. The stochastic nature of the tutor firing also led to a decrease in learning accuracy as the tutor error integration timescale τtutor increased (Figure 5D). This happens through two related effects: (1) the signal-to-noise ratio in the tutor guiding signal decreases as τtutor increases once the tutor error integration timescale is longer than the duration T of the motor program (see Appendix); and (2) the fluctuations in the conductor -- student weights lead to some weights getting clamped at 0 due to the positivity constraint, which leads to the motor program overshooting the target (see Appendix). The latter effect can be reduced by either allowing for negative weights, or changing the motor output to a push-pull architecture in which some student neurons enhance the output while others inhibit it. The signal-to-noise ratio effect can be attenuated by increasing the gain of the tutor signal, which inhibits early learning, but improves the quality of the guiding signal in the latter stages of the learning process. It is also worth emphasizing that these effects only become relevant once the tutor error integration timescale τtutor becomes significantly longer than the duration of the motor program, T , which for a birdsong motif would be around 1 second. Spiking in our model tends to be a little more regular than that in the recordings (compare Figure 5A 9 Figure 5: Results from simulations in spiking neural networks. A. Spike patterns recorded from zebra finch RA during song production, for a juvenile (top) and an adult (bottom). Each color corresponds to a single neuron, and the song-aligned spikes for six renditions of the song are shown. Adapted from ( Olveczky, Otchy, et al. 2011). B. Spike patterns from model student neurons in our simulations, for the untrained (top) and trained (bottom) models. The training used α = 1, β = 0, and τtutor = 80 ms, and ran for 600 iterations of the song. Each model neuron corresponds to a different output channel of the simulation. In this case, the targets for each channel were chosen to roughly approximate the time course observed in the neural recordings. C. Progression of reproduction error in the spiking simulation as a function of the number of repetitions for the same conditions as in panel B. The inset shows the accuracy of reproduction in the trained model for one of the output channels. (See online Video 5.) D. Effects of mismatch between student and tutor on reproduction accuracy in the spiking model. The heatmap shows the final reproduction error of the motor output after 1000 learning cycles in a spiking simulation where a student with parameters α, β, τ1, and τ2 was paired with a tutor with memory timescale τtutor. On the y axis, τ1 and τ2 were kept fixed at 80 ms and 40 ms, respectively, while α and β were varied (subject to the constraint α − β = 1; see section 2.2). Different choices of α and β lead to different optimal timescales τ∗ tutor according to eq. (4). The diagonal elements correspond to matched tutor and student, τtutor = τ∗ tutor. Note that the color scale is logarithmic. Video 5: Evolution of motor output during learning in a spiking simulation. The plasticity rule parameters were α = 1, β = 0, and the tutor had a matching timescale τtutor = τ∗ tutor = 80 ms. This video relates to Figure 5C. 10 102040801603206401280256051201024020480τtutor102040801603206401280256051201024020480τ∗tutor=ατ1−βτ2α−β0.501.002.005.0010.000100200300400500600repetitions02468101214errorα=1.0,β=0.0,τtutor=80.0time0102030405060708090outputtargetoutput06000100200300400500600t(ms)0100200300400500600juvenileadultjuvenileadultABCD01002003004005000100200300400500t(ms)worsebetter and Figure 5B). This could be due to sources of noise that are present in the brain which we did not model. One detail that our model does not capture is the fact that many LMAN spikes occur in bursts, while in our simulation LMAN firing is Poisson. Bursts are more likely to produce spikes in downstream RA neurons particularly because of the NMDA dynamics, and thus a bursty LMAN will be more effective at injecting variability into RA (Kojima, Kao, and Doupe 2013). Small inaccuracies in aligning the recorded spikes to the song are also likely to contribute apparent variability between renditions in the experiment. Indeed, some of the variability in Figure 5A looks like it could be due to time warping and global time shifts that were not fully corrected. 2.6 Robust learning with credit assignment errors The calculation of the tutor output in our rule involved estimating the motor error j from eq. (2). This required knowledge of the assignment between student activities and motor output, which in our model was represented by the matrix Maj (Figure 2A). In our simulations, we typically chose an assignment in which each student neuron contributed to a single output channel, mimicking the empirical findings for neurons in bird RA. Mathematically, this implies that each column of Maj contained a single non-zero element. In Figure 6A, we show what happened in the rate-based model when the tutor incorrectly assigned a certain fraction of the neurons to the wrong output. Specifically, we considered two output channels, y1 and y2, with half of the student neurons contributing only to y1 and the other half contributing only to y2. We then scrambled a fraction ρ of this assignment when calculating the motor error, so that the tutor effectively had an imperfect knowledge of the student -- output relation. Figure 6A shows that learning is robust to this kind of mis-assignment even for fairly large values of the error fraction ρ up to about 40%, but quickly deteriorates as this fraction approaches 50%. Due to environmental factors that affect development of different individuals in different ways, it is unlikely that the student -- output mapping can be innate. As such, the tutor circuit must learn the mapping. Indeed, it is known that LMAN in the bird receives an indirect evaluation signal via Area X, which might be used to effect this learning (Andalman and Fee 2009; Gadagkar et al. 2016; Hoffmann et al. 2016; Kubikova and Kost´al 2010). One way in which this can be achieved is through a reinforcement paradigm. We thus considered a learning rule where the tutor circuit receives a reward signal that enables it to infer the student -- output mapping. In general the output of the tutor circuit should depend on an integral of the motor error, as in eq. (3), to best instruct the student. For simplicity, we start with the memory-less case, τtutor = 0, in which only the instantaneous value of the motor error is reflected in the tutor signal; we then show how to generalize this for τtutor > 0. As before, we took the tutor neurons to fire Poisson spikes with time-dependent rates fj(t), which were initialized arbitrarily. Because of stochastic fluctuations, the actual tutor activity on any given trial, gj(t), differs somewhat from the average, ¯gj(t). Denoting the difference by ξj(t) = gj(t) − ¯gj(t), the update rule for the tutor firing rates was given by (6) where ηtutor is a learning rate, R(t) is the instantaneous reward signal, and ¯R is its average over recent renditions of the motor program. In our implementation, ¯R is obtained by convolving R(t) with an exponential kernel (timescale = 1 second). The reward R(tmax) at the end of one rendition becomes the baseline at the start of the next rendition R(0). The baseline ¯gj(t) of the tutor activity is calculated by averaging over recent renditions of the song with exponentially decaying weights (one e-fold of decay for every 5 renditions). Further implementation details are available in our code at https://github.com/ttesileanu/twostagelearning. ∆fj(t) = ηtutor(R(t) − ¯R)ξj(t) , The intuition behind this rule is that, whenever a fluctuation in the tutor activity leads to better-than- average reward (R(t) > ¯R), the tutor firing rate changes in the direction of the fluctuation for subsequent trials, "freezing in" the improvement. Conversely, the firing rate moves away from the directions in which fluctuations tend to reduce the reward. To test our learning rule, we ran simulations using this reinforcement strategy and found that learning again converges to an accurate rendition of the target output (Figure 6B, inset and online Video 6). The number of repetitions needed for training is greatly increased compared to the case in which the credit assignment is assumed known by the tutor circuit (compare Figure 6B to Figure 5C). This is because the tutor needs to use many training rounds for experimentation before it can guide the conductor -- student plasticity. The rate of learning in our model is similar to the songbird (i.e., order 10 000 repetitions for learning, given that a zebra 11 Figure 6: Credit assignment and reinforcement learning. A. Effects of credit mis-assignment on learning in a rate-based simulation. Here, the system learned output sequences for two independent channels. The student -- output weights Maj were chosen so that the tutor wrongly assigned a fraction of student neurons to an output channel different from the one it actually mapped to. The graph shows how the accuracy of the motor output after 1000 learning steps depended on the fraction of mis-assigned credit. B. Learning curve and trained motor output (inset) for one of the channels showing two-stage reinforcement-based learning for the memory-less tutor (τtutor = 0). The accuracy of the trained model is as good as in the case where the tutor was assumed to have a perfect model of the student -- output relation. However, the speed of learning is reduced. (See online Video 6.) C. Learning curve and trained motor output (inset) for one of the output channels showing two-stage reinforcement-based learning when the tutor circuit needs to integrate information about the motor error on a certain timescale. Again, learning was slow, but the accuracy of the trained state was unchanged. (See online Video 7.) D. Evolution of the average number of HVC inputs per RA neuron with learning in a reinforcement example. Synapses were considered pruned if they admitted a current smaller than 1 nA after a pre-synaptic spike in our simulations. Video 6: Evolution of motor output during learning in a spiking simulation with a reinforcement-based tutor. Here the tutor was memory-less (τtutor = 0). This video relates to Figure 6B. Video 7: Evolution of motor output during learning in a spiking simulation with a reinforcement-based tutor. Here the tutor needed to integrate information about the motor error on a timescale τtutor = 440 ms. This video relates to Figure 6C. 12 0200040006000800010000repetitions050100150200conductor inputsperstudent neuron0.00.10.20.30.40.5fractionmismatch0246810finalerror02000400060008000repetitions02468101214errorα=10.0,β=9.0,τtutor=440.0time020406080outputtargetoutput02000400060008000repetitions02468101214errorα=1.0,β=0.0,τtutor=0.0time020406080outputtargetoutput0600ABCD0600 finch typically sings about 1000 repetitions of its song each day, and takes about one month to fully develop adult song). Because of the extra training time needed for the tutor to adapt its signal, the motor output in our reward-based simulations tends to initially overshoot the target (leading to the kink in the error at around 2000 repetitions in Figure 6B). Interestingly, the subsequent reduction in output that leads to convergence of the motor program, combined with the positivity constraint on the synaptic strengths, leads to many conductor -- student connections being pruned (Figure 6D). This mirrors experiments on songbirds, where the number of connections between HVC and RA first increases with learning and then decreases (Garst-Orozco, Babadi, and Olveczky 2014). The reinforcement rule described above responds only to instantaneous values of the reward signal and tutor firing rate fluctuations. In general, effective learning requires that the tutor keep a memory trace of its activity over a timescale τtutor > 0, as in eq. (4). To achieve this in the reinforcement paradigm, we can use a simple generalization of eq. (6) where the update rule is filtered over the tutor memory timescale: (cid:90) t ∆fj(t) = ηtutor 1 τtutor (cid:48) dt (R(t (cid:48) (cid:48) ) − ¯R)ξj(t )e −(t−t(cid:48) )/τtutor . (7) We tested that this rule leads to effective learning when paired with the corresponding student, i.e., one for which eq. (4) is obeyed (Figure 6C and online Video 7). The reinforcement rules proposed here are related to the learning rules from (Fiete, Fee, and Seung 2007; Fiete and Seung 2006) and (Farries and Fairhall 2007). However, those models focused on learning in a single pass, instead of the two-stage architecture that we studied. In particular, in (Fiete, Fee, and Seung 2007), area LMAN was assumed to generate pure Poisson noise and reinforcement learning took place at the HVC -- RA synapses. In our model, which is in better agreement with recent evidence regarding the roles of RA and LMAN in birdsong (Andalman and Fee 2009), reinforcement learning first takes place in the anterior forebrain pathway (AFP), for which LMAN is the output. A reward-independent heterosynaptic plasticity rule then solidifies the information in RA. In our simulations, tutor neurons fire Poisson spikes with specific time-dependent rates which change during learning. The timecourse of the firing rates in each repetition must then be stored somewhere in the brain. In fact, in the songbird, there are indirect projections from HVC to LMAN, going through the basal ganglia (Area X) and the dorso-lateral division of the medial thalamus (DLM) in the anterior forebrain pathway (Figure 1A) (Perkel 2004). These synapses could store the required time-dependence of the tutor firing rates. In addition, the same synapses can provide the timebase input that would ensure synchrony between LMAN firing and RA output, as necessary for learning. Our reinforcement learning rule for the tutor area, eq. (6), can be viewed as an effective model for plasticity in the projections between HVC, Area X, DLM, and LMAN, as in (Fee and Goldberg 2011). In this picture, the indirect HVC -- LMAN connections behave somewhat like the "hedonistic synapses" from (Seung 2003), though we use a simpler synaptic model here. Implementing the integral from eq. (7) would require further recurrent circuitry in LMAN which is beyond the scope of this paper, but would be interesting to investigate in future work. 3 Discussion We built a two-stage model of learning in which one area (the student) learns to perform a sequence of actions under guidance from a tutor area. This architecture is inspired by the song system of zebra finches, where area LMAN provides a corrective bias to the song that is then consolidated in the HVC -- RA synapses. Using an approach rooted in the efficient coding literature, we showed analytically that, in a simple model, the tutor output that is most likely to lead to effective learning by the student involves an integral over the recent magnitude of the motor error. We found that efficiency requires that the timescale for this integral should be related to the synaptic plasticity rule used by the student. Using simulations, we tested our findings in more general settings. In particular, we demonstrated that tutor-student matching is important for learning in a spiking-neuron model constructed to reproduce spiking patterns similar to those measured in zebra finches. Learning in this model changes the spiking statistics of student neurons in realistic ways, for example, by producing more bursty, stereotyped firing events as learning progresses. Finally, we showed how the tutor can build its error-correcting signal by means of reinforcement learning. 13 If the birdsong system supports efficient learning, our model can predict the temporal structure of the firing patterns of RA-projecting LMAN neurons, given the plasticity rule implemented at the HVC -- RA synapses. These predictions can be directly tested by recordings from LMAN neurons in singing birds, assuming that a good measure of motor error is available, and that we can estimate how the neurons contribute to this error. Moreover, recordings from a tutor circuit, such as LMAN, could be combined with a measure of motor error to infer the plasticity rule in a downstream student circuit, such as RA. This could be compared with direct measurements of the plasticity rule obtained in slice. Conversely, knowledge of the student plasticity rule could be used to predict the time-dependence of tutor firing rates. According to our model, the firing rate should reflect the integral of the motor error with the timescale predicted by the model. A different approach would be to artificially tutor RA by stimulating LMAN neurons electrically or optogenetically. We would predict that if the tutor signal is delivered appropriately (e.g., in conjunction with a particular syllable (Tumer and Brainard 2007)), then the premotor bias produced by the stimulation should become incorporated into the motor pathway faster when the timescale of the artificial LMAN signal is properly matched to the RA synaptic plasticity rule. Our model can be applied more generally to other systems in the brain exhibiting two-stage learning, such as motor learning in mammals. If the plasticity mechanisms in these systems are different from those in songbirds, our predictions for the structure of the guiding signal will vary correspondingly. This would allow a further test of our model of "efficient learning" in the brain. It is worth pointing out that our model was derived assuming a certain hierarchy among the timescales that model the student plasticity and the tutor signal. A mismatch between the model predictions and observations could also imply a breakdown of these approximations, rather than failure of the hypothesis that the particular system under study evolved to support efficient learning. Of course our analysis could be extended by relaxing these assumptions, for example by keeping more terms in the Taylor expansion that we used in our derivation of the matched tutor signal. Applied to birdsong, our model is best seen as a mechanism for learning song syllables. The ordering of syllables in song motifs seems to have a second level of control within HVC and perhaps beyond (Basista et al. 2014; Hamaguchi, Tanaka, and Mooney 2016). Songs can also be distorted by warping their timebase through changes in HVC firing without alterations of the HVC -- RA connectivity (Ali et al. 2013). In view of these phenomena, it would be interesting to incorporate our model into a larger hierarchical framework in which the sequencing and temporal structure of the syllables are also learned. A model of transitions between syllables can be found in (Doya and Sejnowski 2000), where the authors use a "weight perturbation" optimization scheme in which each HVC -- RA synaptic weight is perturbed individually. We did not follow this approach because there is no plausible mechanism for LMAN to provide separate guidance to each HVC -- RA synapse; in particular, there are not enough LMAN neurons (Fiete, Fee, and Seung 2007). In this paper we assumed a two-stage architecture for learning, inspired by birdsong. An interesting question is whether and under what conditions such an architecture is more effective than a single-step model. Possibly, having two stages is better when a single tutor area is responsible for training several different dedicated controllers, as is likely the case in motor learning. It would then be beneficial to have an area that can learn arbitrary behaviors, perhaps at the cost of using more resources and having slower reaction times, along with the ability to transfer these behaviors into low-level circuitry that is only capable of producing stereotyped motor programs. The question then arises whether having more than two levels in this hierarchy could be useful, what the other levels might do, and whether such hierarchical learning systems are implemented in the brain. Acknowledgments We would like to thank Serena Bradde for fruitful discussions during the early stages of this work. We also thank Xuexin Wei and Christopher Glaze for useful discussions. We are grateful to Timothy Otchy for providing us with some of the data we used in this paper. During this work VB was supported by NSF grant PHY-1066293 at the Aspen Center for Physics and by NSF Physics of Living Systems grant PHY-1058202. TT was supported by the Swartz Foundation. 14 A Methods A.1 Equations for rate-based model The basic equations we used for describing our rate-based model (Figure 2A) are the following: ya(t) = Majsj(t) , sj(t) = Wijci(t) + wgj(t) − xinh . j (cid:88) (cid:88) (cid:88) i (cid:90) t (A.1) (A.2) In simulations, we further filtered the output using an exponential kernel, ya(t) = Maj (cid:48) sj(t ) e −(t−t(cid:48) (cid:48) )/τout dt , j 0 with a timescale τout that we typically set to 25 ms. The smoothing produces more realistic outputs by mimicking the relatively slow reaction time of real muscles, and stabilizes learning by filtering out high- frequency components of the motor output. The latter interfere with learning because of the delay between the effect of conductor activity on synaptic strengths vs. motor output. This delay is of the order τ1,2 − τout (see the plasticity rule below). The conductor activity in the rate-based model is modeled after songbird HVC (Hahnloser, Kozhevnikov, and Fee 2002): each neuron fires a single burst during the motor program. Each burst corresponds to a sharp increase of the firing rate ci(t) from 0 to a constant value, and then a decrease 10 ms later. The activities of the different neurons are spread out to tile the whole duration of the output program. Other choices for the conductor activity also work, provided no patterns are repeated (see Appendix). A.2 Mathematical description of plasticity rule In our model the rate of change of the synaptic weights obeys a rule that depends on a filtered version of the conductor signal (see Figure 2B). This is expressed mathematically as where η is a learning rate and ci = K ∗ ci, with the star representing convolution and K being a filtering kernel. We considered a linear combination of two exponential kernels with timescales τ1 and τ2, dWij dt = η ci(t) (gj(t) − θ) , (A.3) with Ki(t) given by Ki(t) = K(t) = αK1(t) − βK2(t) , (cid:40) τ −1 i e−t/τi 0 for t ≥ 0, else. (A.4) (A.5) Different choices for the kernels give similar results (see Appendix). The overall scale of α and β can be absorbed into the learning rate η in eq. (A.3). In our simulations, we fix α − β = 1 and keep the learning rate constant as we change the plasticity rule (see eq. (3)). In the spiking simulations with and without reinforcement learning in the tutor circuit, the firing rates ci(t) and gj(t) were estimated by filtering spike trains with exponential kernels whose timescales were in the range 5 ms -- 40 ms. The reinforcement studies typically required longer timescales for stability, possibly because of delays between conductor activity and reward signals. A.3 Derivation of the matching tutor signal To find the tutor signal that provides the most effective teaching for the student, we first calculate how much synaptic weights change according to our plasticity rule, eq. (A.3). Then we require that this change matches the gradient descent direction. We have ci(t)(gj(t) − θ) dt . (A.6) (cid:90) T ∆Wij = 0 dWij dt (cid:90) T dt = η 0 15 Because of the linearity assumptions in our model, it is sufficient to focus on a case in which each conductor neuron, i, fires a single short burst, at a time ti. We write this as ci(t) = δ(t − ti), and so (cid:90) T (cid:90) T ∆Wij = 0 dWij dt dt = η 0 K(t − ti)(gj(t) − θ) dt , (A.7) where we used the definition of ci(t). If the time constants τ1, τ2 are short compared to the timescale on which the tutor input gj(t) varies, only the values of gj(t) around time ti will contribute to the integral. If we further assume that T (cid:29) ti, we can use a Taylor expansion of gj(t) around t = ti to perform the calculation: (cid:90) ∞ ∆Wij ≈ η ti = η(gj(ti) − θ) = η(gj(ti) − θ) (cid:48) K(t) dt + ηg K(t − ti)(cid:0)gj(ti) − θ + (t − ti)g j(ti)(cid:1) dt (cid:90) ∞ (cid:0)αK1(t) − βK2(t)(cid:1) dt + ηg (cid:90) ∞ (cid:90) ∞ (cid:90) ∞ ∆Wij = η(cid:2)(α − β) (gj(ti) − θ) + (ατ1 − βτ2)g (cid:48) j(ti) (cid:48) j(ti) tK(t) dt 0 0 0 0 Doing the integrals involving the exponential kernels K1 and K2, we get t(cid:0)αK1(t) − βK2(t)(cid:1) dt . j(ti)(cid:3) . (cid:48) (A.8) (A.9) We would like this synaptic change to optimally reduce a measure of mismatch between the output and the desired target as measured by a loss function. A generic smooth loss function L(ya(t), ¯ya(t)) can be quadratically approximated when ya is sufficiently close to the target ¯ya(t). With this in mind, we consider a quadratic loss (cid:90) T (cid:88) a 0 (cid:2)ya(t) − ¯ya(t)(cid:3)2 L = 1 2 dt . (A.10) The loss function would decrease monotonically during learning if synaptic weights changed in proportion to the negative gradient of L: where γ is a learning rate. This implies ∆Wij = −γ ∆Wij = −γ ∂L ∂Wij , (cid:90) T (cid:88) a 0 (cid:2)ya(t) − ¯ya(t)(cid:3) ci(t) . Maj (A.11) (A.12) (A.13) Using again ci(t) = δ(t − ti), we obtain ∆Wij = −γj(ti) , where we used the notation from eq. (2) for the motor error at student neuron j. We now set (A.9) and (A.13) equal to each other. If the conductor fires densely in time, we need the equality to hold for all times, and we thus get a differential equation for the tutor signal gj(t). This identifies the tutor signal that leads to gradient descent learning as a function of the motor error j(t), eq. (3) (with the notation ζ = γ/η). A.4 Spiking simulations We used spiking models that were based on leaky integrate-and-fire neurons with current-based dynamics for the synaptic inputs. The magnitude of synaptic potentials generated by the conductor -- student synapses was independent of the membrane potential, approximating AMPA receptor dynamics, while the synaptic inputs from the tutor to the student were based on a mixture of AMPA and NMDA dynamics. Specifically, the 16 equations describing the dynamics of the spiking model were: (cid:1) = (VR − Vj) + R(cid:0)I AMPA (cid:88) − Vinh , δ(t − tconductor #i (cid:88) I AMPA j τAMPA (cid:88) + I NMDA = − Wij + k k i j j δ(t − ttutor k ) , τm dVj dt dI AMPA j dt dI NMDA j dt (cid:88) k ) + (1 − r)w + rwG(Vj) k Sj(t) , = − I NMDA j τNMDA ginh Nstudent Sj τinh + (cid:88) (cid:88) j k Vinh = dSj dt = − (cid:20) δ(t − tstudent k ) , (cid:21)−1 G(V ) = 1 + [Mg] 3.57 mM exp(−V /16.13 mV) . (except during refractory period) δ(t − ttutor k ) , (A.14) Here Vj is the membrane potential of the jth student neuron and VR is the resting potential, as well as the potential to which the membrane was reset after a spike. Spikes were registered whenever the membrane potential went above a threshold Vth, after which a refractory period τref ensued. Apart from excitatory AMPA and NMDA inputs modeled by the I AMPA variables in our model, we also included a global inhibitory signal Vinh which is proportional to the overall activity of student neurons averaged over a timescale τinh. The averaging is performed using the auxiliary variables Sj which are convolutions of student spike trains with an exponential kernel. These can be thought of as a simple model for the activities of inhibitory interneurons in the student. and I NMDA j j Table 1 gives the values of the parameters we used in the simulations. These values were chosen to match the firing statistics of neurons in bird RA, as described below. Parameter Symbol Value Parameter Symbol Value No. of conductor neurons Reset potential Threshold potential Membrane time constant Refractory period VR Vth τm τref AMPA time constant τAMPA 6.3 ms 300 No. of student neurons −72.3 mV Input resistance −48.6 mV Strength of inhibition 24.5 ms 1.1 ms Fraction NMDA receptors Strength of synapses from tutor No. of conductor synapses per student neuron NMDA time constant τNMDA 81.5 ms Mean strength of synapses Time constant for global in- hibition Conductor firing rate during bursts τinh 20 ms 632 Hz from conductor Standard of conductor -- student weights deviation Table 1: Values for parameters used in the spiking simulations. R ginh r w 80 353 MΩ 1.80 mV 0.9 100 nA 148 32.6 nA 17.4 nA The voltage dynamics for conductor and tutor neurons was not simulated explicitly. Instead, each conductor neuron was assumed to fire a burst at a fixed time during the simulation. The onset of each burst had additive timing jitter of ±0.3 ms and each spike in the burst had a jitter of ±0.2 ms. This modeled the uncertainty in spike times that is observed in in vivo recordings in birdsong (Hahnloser, Kozhevnikov, and Fee 2002). Tutor neurons fired Poisson spikes with a time-dependent firing rate that was set as described in the main text. The initial connectivity between conductor and student neurons was chosen to be sparse (see Table 1). The initial distribution of synaptic weights was log-normal, matching experimentally measured values for zebra finches (Garst-Orozco, Babadi, and Olveczky 2014). Since these measurements are done in the slice, the absolute number of HVC synapses per RA neuron is likely to have been underestimated. The number 17 of conductor -- student synapses we start with in our simulations is thus chosen to be higher than the value reported in that paper (see Table 1), and is allowed to change during learning. We checked that the learning paradigm described here is robust to substantial changes in these parameters, but we have chosen values that are faithful to birdsong experiments and which are thus able to imitate the RA spiking statistics during song. The synapses projecting onto each student neuron from the tutor have a weight that is fixed during our simulations reflecting the finding in (Garst-Orozco, Babadi, and Olveczky 2014) that the average strength of LMAN -- RA synapses for zebra finches does not change with age. There is some evidence that individual LMAN -- RA synapses undergo plasticity concurrently with the HVC -- RA synapses (Mehaffey and Doupe 2015) but we did not seek to model this effect. There are also developmental changes in the kinetics of NMDA-mediated synaptic currents in both HVC -- RA and LMAN -- RA synapses which we do not model (Stark and Perkel 1999). These, however, happen early in development, and thus are unlikely to have an effect on song crystallization, which is what our model focuses on. Stark and Perkel (1999) also observed changes in the relative contribution of NMDA to AMPA responses in the HVC -- RA synapses. We do not incorporate such effects in our model since we do not explicitly model the dynamics of HVC neurons in this paper. However, this is an interesting avenue for future work, especially since there is evidence that area HVC can also contribute to learning, in particular in relation to the temporal structure of song (Ali et al. 2013). A.5 Matching spiking statistics with experimental data We used an optimization technique to choose parameters to maximize the similarity between the statistics of spiking in our simulations and the firing statistics observed in neural recordings from the songbird. The comparison was based on several descriptive statistics: the average firing rate; the coefficient of variation and skewness of the distribution of inter-spike intervals; the frequency and average duration of bursts; and the firing rate during bursts. For calculating these statistics, bursts were defined to start if the firing rate went above 80 Hz and last until the rate decreased below 40 Hz. To carry out such optimizations in the stochastic context of our simulations, we used an evolutionary algorithm -- the covariance matrix adaptation evolution strategy (CMA-ES) (Hansen 2006). The objective and the observed statistics xobs function was based on the relative error between the simulation statistics xsim , i i (cid:34)(cid:88) (cid:18) xsim i i − 1 xobs i (cid:19)2(cid:35)1/2 error = . (A.15) Equal weight was placed on optimizing the firing statistics in the juvenile (based on a recording from a 43 dph bird) and optimizing firing in the adult (based on a recording from a 160 dph bird). In this optimization there was no learning between the juvenile and adult stages. We simply required that the number of HVC synapses per RA neuron, and the mean and standard deviation of the corresponding synaptic weights were in the ranges seen in the juvenile and adult by Garst-Orozco, Babadi, and Olveczky (2014). The optimization was carried out in Python (RRID:SCR_008394), using code from https://www.lri.fr/~hansen/cmaes_inmatlab.html. The results fixed the parameter choices in Table 1 which were then used to study our learning paradigm. While these choices are important for achieving firing statistics that are similar to those seen in recordings from the bird, our learning paradigm is robust to large variations in the parameters in Table 1. A.6 Software and data We used custom-built Python (RRID:SCR_008394) code for simulations and data analysis. The software and data that we used can be accessed online on GitHub (RRID:SCR_002630) at https://github.com/ttesileanu/ twostagelearning. 18 Suppose further that we use a general loss function, (B.1) (B.2) B Appendix B.1 Effect of nonlinearities We can generalize the model from eq. (A.1) by using a nonlinear transfer function from student activities to motor output, and a nonlinear activation function for student neurons: (cid:17) (cid:17) . (cid:1) dt . (cid:12)(cid:12)(cid:12)(cid:12)ya(t)−¯ya(t) ∂L ∂ya j (cid:16)(cid:88) (cid:16)(cid:88) (cid:90) T i (cid:0) (cid:90) T (cid:88) 0 a (cid:88) a ya(t) = Na Majsj(t) , sj(t) = F Wijci(t) + wgj(t) − xinh L = 0 L {ya(t) − ¯ya(t)} ∆Wij = −γ MajN (cid:48) aF (cid:48) ci(t) (cid:80) j(t) = MajN (cid:48) a ∂L ∂ya (cid:12)(cid:12)(cid:12)(cid:12)ya(t)−¯ya(t) Carrying out the same argument as that from section A.3, the gradient descent condition, eq. (A.11), implies The departure from the quadratic loss function, L (cid:54)= 1 Na, have the effect of redefining the motor error, 2 a(ya(t) − ¯ya(t))2, and the nonlinearities in the output, . (B.3) . (B.4) A proper loss function will be such that the derivatives ∂L/∂ya vanish when ya(t) = ¯ya(t), and so the motor error j as defined here is zero when the rendition is perfect, as expected. If we use a tutor that ignores the nonlinearities in a nonlinear system, i.e., if we use eq. (2) instead of eq. (B.4) to calculate the tutor signal that is plugged into eq. (3), we still expect successful learning provided that N(cid:48) a > 0 and that L is itself an increasing function of ya − ¯ya (see section B.2). This is because replacing eq. (B.4) with eq. (2) would affect the magnitude of the motor error without significantly changing its direction. In more complicated scenarios, if the transfer function to the output is not monotonic, there is the potential that using eq. (2) would push the system away from convergence instead of towards it. In such a case, an adaptive mechanism, such as the reinforcement rules from eqns. (6) or (7) can be used to adapt to the local values of the derivatives N(cid:48) a and ∂L/∂ya. Finally, the nonlinear activation function F introduces a dependence on the student output sj(t) in eq. (B.3), since F (cid:48) is evaluated at F −1(sj(t)). To obtain a good match between the student and the tutor in this context, we can modify the student plasticity rule (A.3) by adding a dependence on the postsynaptic activity, dWij dt = η ci(t) (gj(t) − θ) F (cid:48) (F −1(sj(t))) . (B.5) In general, synaptic plasticity has been observed to indeed depend on postsynaptic activity (Chistiakova, Bannon, et al. 2014; Chistiakova and Volgushev 2009). Our derivation suggests that the effectiveness of learning could be improved by tuning this dependence of synaptic change on postsynaptic activity to the activation function of postsynaptic neurons, according to eq. (B.5). It would be interesting to check whether such tuning occurs in real neurons. B.2 Effect of different output functions In the main text, we assumed a linear mapping between student activities and motor output. Moreover, we assumed a myotopic organization, in which each student neuron projected to a single muscle, leading to a student -- output assignment matrix Maj in which each column had a single non-zero entry. We also assumed 19 that student neurons only contributed additively to the outputs, with no inhibitory activity. Here we show that our results hold for other choices of student -- output mappings. For example, assume a push-pull architecture, in which half of the student neurons controlling one output are excitatory and half are inhibitory. This can be used to decouple the overall firing rate in the student from the magnitude of the outputs. Learning works just as effectively as in the case of the purely additive student -- output mapping when using matched tutors, Appendix Figures 1A and 1B. The consequences of mismatching student and tutor circuits are also not significantly changed, Appendix Figures 1C and 1D. We can also consider nonlinear mappings between the student activity and the final output. If there is a monotonic output nonlinearity, as in eq. (B.1) with N(cid:48) a > 0, the tutor signal derived for the linear case, eq. (3), can still achieve convergence, though at a slower rate and with a somewhat lower accuracy (see Appendix Figure 1E for the case of a sigmoidal nonlinearity). For non-monotonic nonlinearities, the direction from which the optimum is approached can be crucial, as learning can get stuck in local minima of the loss function.2 Studying this might provide an interesting avenue to test whether learning in songbirds is based on a gradient descent-type rule or on a more sophisticated optimization technique. B.3 Different inhibition models In the spiking model, we used an activity-dependent inhibitory signal that was proportional to the average student activity. Using a constant inhibition instead, Vinh = constant, does not significantly change the results: see Appendix Figure 1F for an example. B.4 Effect of changing plasticity kernels In the main text, we used exponential kernels with τ1 = 80 ms and τ2 = 40 ms for the smoothing of the conductor signal that enters the synaptic plasticity rule, eq. (A.3). We can generalize this in two ways: we can use different timescales τ1, τ2, or we can use a different functional form for the kernels. (Note that in the main text we showed the effects of varying the parameters α and β in the plasticity rule, while the timescales τ1 and τ2 were kept fixed.) The values for the timescales τ1,2 were chosen to roughly match the shape of the plasticity curve measured in slices of zebra finch RA (Mehaffey and Doupe 2015) (see Figures 1C, 1D). The main predictions of our model, that learning is most effective when the tutor signal is matched to the student plasticity rule, and that large mismatches between tutor and student lead to impaired learning, hold well when the student timescales change: see Appendix Figure 2A for the case when τ1 = 20 ms and τ2 = 10 ms. In the main text we saw that the negative effects of tutor -- student mismatch diminish for timescales that are shorter than ∼ τ1,2. In Appendix Figure 2A, the range of timescales where a precise matching is not essential becomes very small because the student timescales are short. Another generalization of our plasticity rule can be obtained by changing the functional form of the kernels used to smooth the conductor input. As an example, suppose K2 is kept exponential, while K1 is replaced by (cid:40) 1 ¯K1(t) = te−t/¯τ1 for t ≥ 0, else. (B.6) ¯τ 2 1 0 An example of learning using an STDP rule based on kernels ¯K1 and K2 where ¯τ1 = τ2 is shown in Appendix Figure 2B. The matching tutor has the same form as before, eq. (3) with timescale τtutor = τ∗ tutor given by eq. (4), but with τ1 = 2¯τ1 = 2τ2. We can see that learning is as effective as in the case of purely exponential kernels. B.5 More general conductor patterns In the main text, we have focused on a conductor whose activity matches that observed in area HVC of songbirds (Hahnloser, Kozhevnikov, and Fee 2002): each neuron fires a single burst during the motor program. Our model, however, is not restricted to this case. We generated alternative conductor patterns by using arbitrarily-placed bursts of activity, as in Appendix Figure 3A. The model converges to a good rendition of the 2We thank Josh Gold for this observation. 20 Appendix figure 1: Robustness of learning. A. Error trace showing how average motor error evolves with repetitions of the motor program for rate-based plasticity paired with a matching tutor, when the student -- output mapping has a push-pull architecture. The inset shows the final motor output (thick red line) compared to the target output (dotted black line). The output on the first rendition and at two other stages of learning are also shown. B. The error trace and final motor output shown for timing-based plasticity matched by a tutor with a long integration timescale. C. Effects of mismatch between student and tutor on reproduction accuracy when using a push-pull architecture for the student -- output mapping. The heatmap shows the final reproduction error of the motor output after 1000 learning cycles when a student with plasticity parameters α and β is paired with a tutor with memory timescale τtutor. Here τ1 = 80 ms and τ2 = 40 ms. D. Error evolution curves as a function of the mismatch between student and tutor. Each plot shows how the error in the motor program changes during 1000 learning cycles for the same conditions as those shown in the heatmap. The region shaded in light pink shows simulations where the mismatch between student and tutor leads to a deteriorating instead of improving performance during learning. E. Convergence in the rate-based model with a linear-nonlinear controller that uses a sigmoidal nonlinearity. F. Convergence in the spiking model when inhibition is constant instead of activity-dependent (Vinh = constant). 21 102040801603206401280256051201024020480τtutor102040801603206401280256051201024020480τ∗tutor=ατ1−βτ2α−β102040801603206401280256051201024020480τtutor102040801603206401280256051201024020480τ∗tutor=ατ1−βτ2α−β0.501.002.005.0010.00050100150200250repetitions02468101214errorα=7.0,β=6.0,τtutor=320.0time01020304050607080outputtargetoutput050100150200250repetitions02468101214errorα=0.0,β=−1.0,τtutor=40.0time010203040506070outputtargetoutput0600ABCD0100200300400500repetitions02468101214errorα=24.0,β=23.0,τtutor=1000.0time01020304050607080outputtargetoutput050100150200repetitions02468101214errorα=1.0,β=0.0,τtutor=80.0time01020304050607080outputtargetoutputEF060006000600worsebetter Appendix figure 2: Effect of changing conductor smoothing kernels in the plasticity rule. (A) Matrix showing learning accuracy when using different timescales for the student plasticity rule. Each entry in the heatmap shows the average rendition error after 1000 learning steps when pairing a tutor with timescale τtutor with a non-matched student. Here the kernels are exponential, with timescales τ1 = 20 ms, τ2 = 10 ms. (B) Evolution of motor error with learning using kernels ∼ e−t/τ and ∼ te−t/τ , instead of the two exponentials used in the main text. The tutor signal is as before, eq. (3). The inset shows the final output for the trained model, for one of the two output channels. Learning is as effective and fast as before. target program, Appendix Figure 3B. Learning is harder in this case because many conductor neurons can be active at the same time, and the weight updates affect not only the output of the system at the current position in the motor program, but also at all the other positions where the conductor neurons fire. This is in contrast to the HVC-like conductor, where each neuron fires at a single point in the motor program, and thus the effect of weight updates is better localized. More generally, simulations show that the sparser the conductor firing, the faster the convergence (data not shown). The accuracy of the final rendition of the motor program (Appendix Figure 3B, inset) is also not as good as before. B.6 Edge effects In our derivation of the matching tutor rule, we assumed that the system has enough time to integrate all the synaptic weight changes from eq. (A.3). However, some of these changes occur tens or hundreds of milliseconds after the inputs that generated them, due to the timescales used in the plasticity kernel. Since our simulations are only run for a finite amount of time, there will in general be edge effects, where periods of the motor program towards the end of the simulations will have difficulty converging. To offset such numerical issues, we ran the simulations for a few hundred milliseconds longer than the duration of the motor program, and ignored the data from this extra period. Our simulations typically run for 600 ms, and the time reserved for relaxation after the end of the program was set to 1200 ms. The long relaxation time was chosen to allow for cases where the tutor was chosen to have a very long memory timescale. B.7 Parameter optimization for reproducing juvenile and adult spiking statistics We set the parameters in our simulations to reproduce spiking statistics from recordings in zebra finch RA as closely as possible. Appendix Figure 4 shows how the distribution of summary statistics obtained from 50 runs of the simulation compares to the distributions calculated from recordings in birds at various developmental stages. Each plot shows a standard box and whisker plot superimposed over a kernel-density estimate of the distribution of a given summary statistic, either over simulation runs or over recordings from birds at various stages of song learning. We ran two sets of simulations, one for a bird with juvenile-like connectivity between HVC and RA, and one with adult-like connectivity (see Methods). In these simulations there was no learning to match the timecourse of songs -- the goal was simply to identify parameters that lead to birdsong-like firing statistics. 22 102040801603206401280256051201024020480τtutor102040801603206401280256051201024020480τ∗tutor=ατ1−βτ2α−β0.501.002.005.0010.00050100150200repetitions02468101214errorα=24.0,β=23.0,τtutor=1000.0time010203040506070outputtargetoutput0600ABworsebetter Appendix figure 3: Learning with arbitrary conductor activity. A. Typical activity of conductor neurons. 20 of the 100 neurons included in the simulation are shown. The activity pattern is chosen so that about 10% of the neurons are active at any given time. The pattern is chosen randomly but is fixed during learning. Each conductor burst lasts 30 ms. B. Convergence curve and final rendition of the motor program (in inset). Learning included two output channels but the final output is shown for only one of them. The qualitative match between our simulations and recordings is good, but the simulations are less variable than the measurements. This may be due to sources of variability that we have ignored -- for example, all our simulated neurons had exactly the same membrane time constants, refractory periods, and threshold potentials, which is not the case for real neurons. Another reason might be that in our simulations, all the runs were performed for the same network, while the measurements are from different cells in different birds. B.8 Effect of spiking stochasticity on learning As pointed out in the main text, learning is affected in the spiking simulations when the tutor error integration timescale τtutor becomes very long. More specifically, two distinct effects occur. First, the fluctuations in the motor output increase, leading to a poorer match to the shape of the target motor program. And second, the whole output gets shifted up, towards higher muscle activation values. Both of these effects can be traced back to the stochasticity of the tutor signal. In the spiking simulations, tutor neurons are assumed to fire Poisson spikes following a time-dependent firing rate that obeys eq. (5). By the nature of the Poisson process, the tutor output in this case will contain fluctuations around the mean, g(t) ∼ ¯g(t) + ξ(t). Recall that the scale of g(t) is set by the threshold θ and thus so is the scale of the variability ξ(t). As long as the tutor error integration timescale is not very long, g(t) roughly corresponds to a smoothed version of the motor error (t) (cf. eq. (5)). However, as τtutor grows past the duration T of the motor program, the exponential term in eq. (5) becomes essentially constant, leading to a tutor signal ¯g(t) whose departures from the center value θ decrease in proportion to the timescale τtutor. As far as the student is concerned, the relevant signal is g(t) − θ (eq. (1)), and thus, when τtutor > T , the signal-to-noise ratio in the tutor guiding signal starts to decrease as 1/τtutor. This ultimately leads to a very noisy rendition of the target program. One way to improve this would be to increase the gain factor ζ that controls the relation between the motor error and the tutor signal (see eq. (5)). This improves the ability of the system to converge onto its target in the late stages of learning. In the early stages of learning, however, this could lead to saturation problems. One way to fix this would be to use a variable gain factor ζ that ensures the whole range of tutor firing rates is used without generating too much saturation. This would be an interesting avenue for future research. Reducing the fluctuations in the tutor signal also decreases the fluctuations in the conductor -- student synaptic weights, which leads to fewer weights being clamped at 0 because of the positivity constraint. This reduces the shift between the learned motor program and the target. As mentioned in the main text, another approach to reducing or eliminating this shift is to allow for negative weights or (more realistically) to use a push-pull mechanism, in which the activity of some student neurons acts to increase muscle output, while the 23 050010001500200025003000repetitions02468101214errorα=5.0,β=4.0,τtutor=240.0time01020304050607080outputtargetoutput06000100200300400500600time(ms)135791113151719conductorneuronindexAB Appendix figure 4: Violin plots showing how the spiking statistics from our simulation compared to the statistics obtained from neural recordings. Each violin shows a kernel-density estimate of the distribution that a particular summary statistic had in either several runs of a simulation, or in several recordings from behaving birds. The circle and the box within each violin show the median and the interquartile range. 24 simjuvenile40-6565-9090-115115-140140-165simadult020406080100120140160Averagefiringrate(Hz)simjuvenile40-6565-9090-115115-140140-165simadult0123456CVofISIsimjuvenile40-6565-9090-115115-140140-165simadult0510152025SkewnessofISIsimjuvenile40-6565-9090-115115-140140-165simadult024681012Burstfrequency(Hz)simjuvenile40-6565-9090-115115-140140-165simadult0.000.050.100.150.20Averageburstlength(s)simjuvenile40-6565-9090-115115-140140-165simadult050100150200250300350400450Firingrateduringbursts(Hz) activity of other student neurons acts as an inhibition on muscle output. C Plasticity parameter values In the heatmaps that appear in many of the figures in the main text and in the supplementary information, we kept the timescales τ1 and τ2 constant while varying α and β to modify the student plasticity rule. Since the overall scale of α and β is inconsequential as it can be absorbed into the learning rate (as explained in section 2.2), we imposed the further constraint α − β = 1. This implies that we effectively focused on a one-parameter family of student plasticity rule, as identified by the value of α (and the corresponding value for β = α − 1). In the figures, we expressed this instead in terms of the timescale of the optimally-matching tutor, τ∗ tutor, as defined in eq. (4). Below we give the explicit values of α and β that we used for each row in the heatmaps. These can be calculated by solving for α in eq. (4), using β = α − 1, and assume that τ1 = 80 ms and τ2 = 40 ms. τ∗ tutor α β 10 −0.75 −1.75 −1.5 20 −1.0 40 0.0 80 2.0 160 6.0 320 640 14.0 30.0 1280 62.0 2560 126.0 5120 10240 254.0 510.0 20480 −0.5 0.0 1.0 3.0 7.0 15.0 31.0 63.0 127.0 255.0 511.0 References [1] Farhan Ali et al. "The basal ganglia is necessary for learning spectral, but not temporal, features of birdsong". In: Neuron 80.2 (Oct. 2013), pp. 494 -- 506. [2] Aaron S. Andalman and Michale S. Fee. "A basal ganglia-forebrain circuit in the songbird biases motor output to avoid vocal errors." In: Proceedings of the National Academy of Sciences of the United States of America 106.30 (July 2009), pp. 12518 -- 23. [3] Mark J. Basista et al. "Independent Premotor Encoding of the Sequence and Structure of Birdsong in Avian Cortex". In: Journal of Neuroscience 34.50 (Dec. 2014), pp. 16821 -- 16834. [4] R. A. Canady et al. "Effect of testosterone on input received by an identified neuron type of the canary song system: a Golgi/electron microscopy/degeneration study". In: The Journal of Neuroscience 8.10 (1988), pp. 3770 -- 3784. [5] Marina Chistiakova, Nicholas M. Bannon, et al. "Heterosynaptic Plasticity: Multiple Mechanisms and Multiple Roles". In: The Neuroscientist : a review journal bringing neurobiology, neurology and psychiatry April (2014). [6] Marina Chistiakova and Maxim Volgushev. "Heterosynaptic plasticity in the neocortex". In: Experimental Brain Research 199.3-4 (2009), pp. 377 -- 390. [7] Kenji Doya and Terrence J. Sejnowski. "A computational model of avian song learning". In: The new cognitive neurosciences. Ed. by Michael S. Gazzaniga. The MIT Press, 2000, pp. 469 -- 482. isbn: 0262071959. arXiv: arXiv:1011.1669v3. [8] Michael A. Farries and Adrienne L. Fairhall. "Reinforcement Learning With Modulated Spike Timing- Dependent Synaptic Plasticity". In: Journal of neurophysiology 98.6 (2007), pp. 3648 -- 3665. [9] Michale S. Fee and Jesse H. Goldberg. "A hypothesis for basal ganglia-dependent reinforcement learning in the songbird". In: Neuroscience 198 (2011), pp. 152 -- 170. 25 [10] [11] Ila R. Fiete, Michale S. Fee, and H. Sebastian Seung. "Model of birdsong learning based on gradient estimation by dynamic perturbation of neural conductances". In: Journal of neurophysiology 98.4 (Oct. 2007), pp. 2038 -- 57. Ila R. Fiete and H. Sebastian Seung. "Gradient learning in spiking neural networks by dynamic per- turbation of conductances". In: Physical Review Letters 97.4 (2006), p. 048104. arXiv: 0601028v1 [arXiv:q-bio]. [12] Paul W. Frankland and Bruno Bontempi. "The organization of recent and remote memories". In: Nature reviews. Neuroscience 6.2 (2005), pp. 119 -- 30. [13] Vikram Gadagkar et al. "Dopamine neurons encode performance error in singing birds". In: Science 354.6317 (2016), pp. 1278 -- 1282. eprint: http://science.sciencemag.org/content/354/6317/1278. full.pdf. [14] Jonathan Garst-Orozco, Baktash Babadi, and Bence P. Olveczky. "A neural circuit mechanism for regulating motor variability during skill learning". In: eLife 3 (2014), e03697. [15] Richard H. R. Hahnloser, Alexay A. Kozhevnikov, and Michale S. Fee. "An ultra-sparse code underlies the generation of neural sequences in a songbird". In: Nature 419.1989 (2002), pp. 796 -- 797. [16] Kosuke Hamaguchi, Masashi Tanaka, and Richard Mooney. "A Distributed Recurrent Network Contributes to Temporally Precise Vocalizations". In: Neuron 91 (2016), pp. 680 -- 693. [17] Nikolaus Hansen. "The CMA Evolution Strategy: A Comparing Review". In: Towards a new evolutionary computation. Advances on estimation of distribution algorithms. 2006, pp. 75 -- 102. [18] Kathrin Herrmann and Arthur P. Arnold. "The development of afferent projections to the robust archistriatal nucleus in male zebra finches: a quantitative electron microscopic study". In: The Journal of Neuroscience 11.7 (1991), pp. 2063 -- 2074. [19] Lukas A. Hoffmann et al. "Dopaminergic Contributions to Vocal Learning". In: The Journal of Neuro- science 36.7 (2016), pp. 2176 -- 2189. [20] K Immelmann. "Song development in the zebra finch and other estrildid finches". In: Bird vocalizations. Ed. by R. A. Hinde. 1969, pp. 61 -- 74. [21] Mimi H. Kao, Brian D. Wright, and Allison J. Doupe. "Neurons in a Forebrain Nucleus Required for Vocal Plasticity Rapidly Switch between Precise Firing and Variable Bursting Depending on Social Context". In: The Journal of Neuroscience 28.49 (2008), pp. 13232 -- 13247. [22] Risa Kawai et al. "Motor Cortex Is Required for Learning but Not for Executing a Motor Skill". In: Neuron 86.3 (2015), pp. 800 -- 812. [23] Satoshi Kojima, Mimi H. Kao, and Allison J. Doupe. "Task-related "cortical" bursting depends critically on basal ganglia input and is linked to vocal plasticity". In: PNAS 110.12 (2013), pp. 4756 -- 4761. [24] L'ubica Kubikova and L'ubor Kost´al. "Dopaminergic system in birdsong learning and maintenance". In: Journal of Chemical Neuroanatomy 39 (2010), pp. 112 -- 123. [25] Anthony Leonardo and Michale S. Fee. "Ensemble Coding of Vocal Control in Birdsong". In: The Journal of Neuroscience 25.3 (2005), pp. 652 -- 661. [26] Galen F. Lynch et al. "Rhythmic Continuous-Time Coding in the Songbird Analog of Vocal Motor Cortex". In: Neuron 90.4 (2016), pp. 877 -- 892. [27] W. Hamish Mehaffey and Allison J. Doupe. "Naturalistic stimulation drives opposing heterosynaptic plasticity at two inputs to songbird cortex". In: Nature Neuroscience 18.9 (2015), pp. 1272 -- 1280. [28] Raoul-Martin Memmesheimer et al. "Learning Precisely Timed Spikes". In: Neuron 82 (2014), pp. 1 -- 14. [29] Bence P. Olveczky, Aaron S. Andalman, and Michale S. Fee. "Vocal experimentation in the juvenile songbird requires a basal ganglia circuit." In: PLoS biology 3.5 (May 2005), e153. [30] Bence P. Olveczky, Timothy M. Otchy, et al. "Changes in the neural control of a complex motor sequence during learning". In: Journal of Neurophysiology 106.1 (2011), pp. 386 -- 397. [31] David J. Perkel. "Origin of the Anterior Forebrain Pathway". In: Annals of the New York Academy of Sciences 1016 (2004), pp. 736 -- 748. 26 [32] Michel A. Picardo et al. "Population-Level Representation of a Temporal Sequence Underlying Song Production in the Zebra Finch". In: Neuron 90.4 (2016), pp. 866 -- 876. [33] H. Sebastian Seung. "Learning in Spiking Neural Networks by Reinforcement of Stochastic Synaptic Transmission". In: 40 (2003), pp. 1063 -- 1073. [34] H. Blair Simpson and David S. Vicario. "Brain Pathways for Learned and Unlearned Vocalizations Differ in Zebra Finches". In: The Journal of Neuroscience 10.5 (1990), pp. 1541 -- 1556. [35] John E. Spiro, Matthew B. Dalva, and Richard Mooney. "Long-Range Inhibition Within the Zebra Finch Song Nucleus RA Can Coordinate the Firing of Multiple Projection Neurons". In: Journal of neurophysiology 81 (1999), pp. 3007 -- 3020. [36] Laura L. Stark and David J. Perkel. "Two-stage, input-specific synaptic maturation in a nucleus essential for vocal production in the zebra finch". In: The Journal of Neuroscience 19.20 (1999), pp. 9107 -- 9116. [37] Evren C. Tumer and Michael S. Brainard. "Performance variability enables adaptive plasticity of 'crystallized' adult birdsong". In: Nature 450.December (2007), pp. 1240 -- 1244. [38] Robert S. Turner and Michel Desmurget. "Basal ganglia contributions to motor control: a vigorous tutor". In: Current Opinion in Neurobiology 20.6 (2010), pp. 704 -- 716. [39] Lena Veit, Dmitriy Aronov, and Michale S. Fee. "Learning to breathe and sing: development of respiratory- vocal coordination in young songbirds". In: Journal of Neurophysiology 106.4 (2011), pp. 1747 -- 1765. eprint: http://jn.physiology.org/content/106/4/1747.full.pdf. [40] Timothy L. Warren et al. "Mechanisms and time course of vocal learning and consolidation in the adult songbird". In: Journal of Neurophysiology 106 (2011), pp. 1806 -- 1821. [41] Albert C. Yu and Daniel Margoliash. "Temporal hierarchical control of singing in birds". In: Science 273.5283 (1996), pp. 1871 -- 1875. 27
1010.4145
2
1010
2010-10-29T14:29:24
Voltage clamp analysis of nonlinear dendritic propertie in prepositus hypoglossi neurons
[ "q-bio.NC" ]
The nonlinear properties of the dendrites in prepositus hypoglossi neurons are involved in maintenance of eye position. The biophysical properties of these neurons are essential for the operation of the vestibular neural integrator that converts a head velocity signal to one that controls eye position. A novel method named QSA (quadratic sinusoidal analysis) for voltage clamped neurons was used to quantify nonlinear responses that are dominated by dendrites. The voltage clamp currents were measured at harmonic and interactive frequencies using specific stimulation frequencies, which act as frequency probes of the intrinsic nonlinear neuronal behavior. These responses to paired frequencies form a matrix that can be reduced by eigendecomposition to provide a very compact piecewise quadratic analysis at different membrane potentials that otherwise is usually described by complex differential equations involving a large numbers of parameters and dendritic compartments. Moreover, the QSA matrix can be interpolated to capture most of the nonlinear neuronal behavior like a Volterra kernel. The interpolated quadratic functions of the two major prepositus hypoglossi neurons, namely type B and D, are strikingly different. A major part of the nonlinear responses is due to the persistent sodium conductance, which appears to be essential for sustained nonlinear effects induced by NMDA activation and thus would be critical for the operation of the neural integrator. Finally, the dominance of the nonlinear responses by the dendrites supports the hypothesis that persistent sodium conductance channels and NMDA receptors act synergistically to dynamically control the influence of individual synaptic inputs on network behavior.
q-bio.NC
q-bio
Voltage clamp analysis of nonlinear dendritic properties in prepositus hypoglossi neurons Christophe Magnani, Daniel Eugène, Erwin Idoux and L.E. Moore* CESEM - UMR8194 - CNRS - Université Paris Descartes Address : 45 rue des Saints-Pères, 75270 PARIS, FRANCE - Tel.:33-(0)142863398, Fax:33-(0)142863399 Keywords: Electrophysiology; Vestibular neural integrator; Membrane potential; Impedance; Quadratic analysis; Persistent sodium conductance; NMDA receptors *Corresponding author Abstract The nonlinear properties of the dendrites in prepositus hypoglossi neurons are in- volved in maintenance of eye position. The biophysical properties of these neurons are essential for the operation of the vestibular neural integrator that converts a head velocity signal to one that controls eye position. A novel method named QSA (quadratic sinusoidal analysis) for voltage clamped neurons was used to quantify nonlinear responses that are dominated by dendrites. The voltage clamp currents were measured at harmonic and interactive frequencies using specific stimulation frequencies, which act as frequency probes of the intrinsic nonlinear neuronal be- havior. These responses to paired frequencies form a matrix that can be reduced by eigendecomposition to provide a very compact piecewise quadratic analysis at different membrane potentials that otherwise is usually described by complex dif- ferential equations involving a large numbers of parameters and dendritic com- partments. Moreover, the QSA matrix can be interpolated to capture most of the nonlinear neuronal behavior like a Volterra kernel. The interpolated quadratic functions of the two major prepositus hypoglossi neurons, namely type B and D, are strikingly different. A major part of the nonlinear responses is due to the per- 1 2 sistent sodium conductance, which appears to be essential for sustained nonlinear effects induced by NMDA activation and thus would be critical for the operation of the neural integrator. Finally, the dominance of the nonlinear responses by the dendrites supports the hypothesis that persistent sodium conductance channels and NMDA receptors act synergistically to dynamically control the influence of individual synaptic inputs on network behavior. Introduction Mathematical models based on the experimentally measured biophysical proper- ties of neurons generally consist of complicated sets of differential equations derived from the historical Hodgkin and Huxley (1952) model. Extending the HH formal- ism to branching neurons requires a large number of parameters that must be determined to obtain a realistic neuronal model. The techniques previously em- ployed to measure these parameters involve either linear admittance (or impedance) measurements or ad hoc extrapolations from voltage clamp experiments with poor space clamp control. Thus, it is important to consider more refined theories from nonlinear analysis, such as nonlinear dynamics of neurons (Gutkin and Ermen- trout, 1998; Izhikevich, 2002) or nonlinear system identification (Marmarelis and Naka, 1973; French, 1976; Victor and Shapley, 1980; Boyd et al., 1983; Schetzen, 2006). A goal of nonlinear analysis is not just a refinement of the linear systems approach, but the development of a fundamental insight into how neurons process information. In a seminal paper, Fitzhugh derived the equations of the nonlinear response for a single sinusoidal voltage clamp (FitzHugh, 1983). This approach has been ex- tended to the quadratic response for a multi-sinusoidal voltage clamp and been developed as a matrix theory termed quadratic sinusoidal analysis (QSA) (Mag- nani and Moore, 2010). The proposed experimental approach is essentially based on QSA, requiring that the stimulus amplitudes evoke mainly linear and quadratic responses. In addition, QSA provides the mathematical tools for a model inde- pendent analysis of quadratic nonlinearities and provides an innovative way to quantitatively describe real neurons and their models. The measurement of non- linearities in neurons under normal physiological conditions is clearly important in order to understand how they process synaptic inputs, which typically evoke 5-10 mV post-synaptic nonlinear responses. Two types of neurons of the rat prepositus hypoglossi nucleus (PHN) were inves- tigated. Both types clearly manifest nonlinearities at multiple subthreshold step levels. The type D neurons are known to show marked spontaneous, voltage depen- dent and irregular oscillatory properties. By contrast, type B neurons, the majority 3 in this nucleus, are non-oscillatory and have regular spontaneous activity that is highly dependent on a significant persistent sodium (gNaP) conductance (Vervaeke et al., 2006). In this paper, the novel QSA method has been used to investigate the quadratic response to time varying voltage clamped stimuli and establish a quan- titative characterization of the nonlinear behavior in order to understand neuronal responses elicited by normal physiological synaptic inputs. It will be shown that at physiological levels of stimulation, neurons and their mod- els can generate significant responses at harmonic and interactive frequencies that are not present in the input signal. Thus, the nonlinear frequency responses contain more frequencies over a wider frequency band than the input signal. As a conse- quence they provide significant amplification at dynamically changing membrane potentials. The use of stimuli with multiple input frequencies allows one to probe neuronal function and characterize it by a matrix of quadratic interactions, namely the QSA matrix. It is then possible to extract information about active membrane properties from this matrix by eigendecomposition. Finally, biologically realistic simulations have been implemented using neuronal models based on vestibular neu- ronal experimental data. These simulations suggest that the nonlinear responses in voltage clamp are dominated by active dendritic structures. Materials and Methods Whole-cell patch-clamp recordings and statistical analysis This paper is both a theoretical and experimental nonlinear approach to neuronal function that adds to previous steady state linear analyses (Fishman et al., 1977; Murphey et al., 1995). It provides a quantitative assessment of quadratic responses of both data recorded from individual neurons and their corresponding biophysical models. Experiments were carried out on male Wistar rats (25- to 52-days-old) supplied by Centre d'Elevage Roger Janvier (Le Genest Saint Isle, France). All ef- forts were made to minimize animal suffering as well as the number of animals used. All experiments followed the guidelines on the ethical use of animals from the Eu- ropean Communities Council Directive of 24 November 1986 (86/609/EEC). Brain dissections were performed as described elsewhere (Idoux et al., 2008). Briefly, af- ter decapitation under deep anesthesia, the brain was quickly removed and placed in ice-cold, phosphate/bicarbonate-buffered artificial cerebro-spinal fluid (ACSF), which included (in mM) 225 sucrose, 5 KCl, 1 NaH2PO4, 26 NaHCO3, 0.25 CaCl2, 1.3 MgCl2, 11 glucose and was bubbled with 95% O2-5% CO2 (pH 7.4). Four or five 250 µm thick, coronal slices containing the PHN were cut from the brainstem with a microslicer (Leica, Rueil-Malmaison, France) and transferred into an incu- 4 bating vial filled with a regular ACSF containing (in mM) 124 NaCl, 5 KCl, 1 NaH2PO4, 26 NaHCO3, 2.5 CaCl2, 1.3 MgCl2, 11 glucose and bubbled with 95% O2 and 5% CO2 (pH 7.4). Slices were then placed one at a time in the recording chamber maintained at 32-34°C, where the slice was superfused with regular ACSF at a constant flow rate of 3 mL min-1. Patch-clamp pipettes were pulled from borosilicate glass tubing to a resistance of 5- 8 MΩ. The control internal solution contained (in mM) 140 K-gluconate, 2 MgCl2, 10 HEPES, 0.1 EGTA, 4 Na2ATP, and 0.4 Na2GTP (adjusted to pH 7.3 with KOH). The junction potential for this internal solution was not subtracted for the potential measurements or the model simulations. PHN neurons were visualized with a Nomarski optic microscope under infrared illumination. Recordings were made with an Axoclamp 2B amplifier (Axon Instruments, Union City, CA, USA) or a Multiclamp 700B (Molecular Devices, Sunnyvale, CA, USA). The spontaneous discharge was first recorded in the current-clamp mode for 8 to 10 minutes once a stable level had been reached and the recorded PHN neuron was determined as B or D type (see Idoux et al., 2008). PHN neurons that had resting membrane potential more negative than -50 mV and a spike amplitude > 45 mV were selected for the voltage clamp experiments. Types D and B neurons from the prepositus hypoglossi nucleus were measured for different stimulus amplitudes and membrane potentials. Based on the criteria for time invariance discussed in the Rationale section, five type D and six type B neurons were selected for detailed analysis. All measurements were made with stimuli applied for twice the duration used in the analysis. Only the last half of the record was used to assure that a steady state condition was reached. At voltage clamp potentials near threshold, transient currents due to uncontrolled action potentials occasionally occurred in the non- analyzed initial part of the recording, however they were completely inactivated by the maintained depolarization with no firing during the latter analyzed part of the record. In some experiments 25-50 µM NMDA (Sigma, St Quentin Fallavier, France) was applied in the presence or absence of 2 µM TTX (Tocris, Bristol, UK). The data acquisition was done with a PC-compatible computer running Win- dows XP, using MATLAB scripts (MATLAB 7.0, MathWorks, Natick, MA, USA). Recordings were low-pass filtered at 2 kHz and digitized at 5 kHz (BNC-2090 + PCI-6052E, National Instruments, Austin, TX, USA). The OneSidedPValue (p- value) of MeanTest[x-y] for paired differences, mean values and standard deviations (±SD) were calculated with the HypothesisTesting package of MATHEMATICA 7.0, (Wolfram Research, Champain, IL, USA). 5 Rationale PHN neurons were analyzed with QSA, which specifically selects harmonic and intermodulation frequencies, described as follows. If a double sinusoidal input has frequencies f1 and f2 then the linear response will have exactly the same frequencies f1 and f2. However, the quadratic response will include additional harmonics 2f1 and 2f2 as well as intermodulation products f1 + f2 and f1 − f2. This principle can be generalized to a multi-sinusoidal input in which case the quadratic response will include double of each input frequency as well as sum and difference of each pair of distinct input frequencies. A quadratic response can generate frequency overlaps when distinct combinations of input frequencies generate the same output frequency. For instance, a multi-sinusoidal input with frequencies 1, 2, 3, 4 (in Hertz) would generate many frequency overlaps such as 1 = 2 − 1 or 2 + 3 = 1 + 4 and so on. In presence of frequency overlaps, it is not possible to unambiguously measure the nonlinear frequency interactions. In the previous example, the measurement at 5 Hz is ambiguous because one is unable to distinguish between the contributions of 2 + 3 and 1 + 4. In order to avoid this problem, the QSA was used with a flexible algorithm generating incommensurable frequencies. This approach is based on a practical measurement technique, namely harmonic probing on Volterra kernels (Boyd et al., 1983; Victor and Shapley, 1980). Since harmonic and intermodulation responses also exist for nonlinearities of higher degrees, for example third order intermodulation products f1 + f2 + f3, it is impor- tant to ensure that the neurons mainly manifest quadratic nonlinearities, otherwise the results would be significantly contaminated. For this, a necessary, but not suf- ficient condition, consists of using relatively small stimulus amplitudes in such a way that only linear and quadratic responses are significant. This approach was used in our previous piecewise linear analysis (Murphey et al., 1995), hence the term piecewise quadratic analysis can be used to describe the QSA extension. The influence of the input amplitude on the harmonic response has been investigated previously (see Moore et al., 1980; FitzHugh, 1983). In order to ensure that the stimulus amplitudes were sufficient to overcome spontaneous noise and avoid signif- icant higher order responses, several algorithms described elsewhere (Magnani and Moore, 2010) have been implemented in MATLAB to verify that the experimental traces are time invariant for both linear and quadratic outputs, and that the signal can be adequately reconstructed by quadratic analysis. It will be shown that the oscillatory type D neurons of the PHN have quadratic responses over a range of subthreshold membrane potentials, namely they con- vert limited amplitude and bandwidth input signals to wider bandwidth and more complex output responses as mentioned above for nonlinear responses. Under nor- mal physiological conditions of current clamp, a depolarization activates gNaP 6 conductance that in turn increases the impedance and consequently reduces the electrotonic length. Type B neurons show similar effects and in addition, they have a significant increase in the electrotonic length at hyperpolarized membrane potentials because of the activation of a hyperpolarization activated conductance (Erchova et al., 2004). Voltage clamp experiments were done to partially control the oscillatory and bistable responses of PHN neurons in order to analyze the nonlinear membrane properties of both the somatic and dendritic regions of these neurons. Indeed, an important advantage of a quantitative voltage clamp analysis of central neurons is the ex- ploitation of the space clamp problem as a way to separate somatic and dendritic responses, namely the voltage clamp current measured from the voltage controlled soma is generally dominated by the unclamped voltage responses of the dendrites (Moore et al., 1999). Thus, this current can be used as a measure of both linear and nonlinear dendritic potential responses while the somatic membrane potential is voltage clamped. These additional currents can flow because of a potential dif- ference between the soma and the rest of the dendrite. In fact, these additional currents reflect the behavior of the dendritic membrane potential that can be taken into account with multi-compartmental models. Previous voltage clamp measurements have been done using signals of small am- plitude to obtain steady state linear responses at different membrane potentials. These measurements were done over a potential range to obtain a quantitative description of the voltage dependent conductances and have allowed the construc- tion of neuronal multi-compartmental models of both type B and D PHN neurons (Idoux et al., 2008). These models allow an estimation of both somatic and den- dritic membrane properties from somatic voltage clamp experiments that probe the soma and all regions of the dendritic structure. When the stimulus amplitudes are sufficiently small to elicit linear responses, both the voltage clamp and the current clamp generate equivalent linear results, however this is not the case for nonlinear responses. The present series of voltage clamp experiments is a quadratic extension of the steady state linear analysis. Voltage clamped neurons show two kinds of nonlinearities : first, space clamped somatic ionic currents, and second, ionic currents in the soma due to an unclamped den- dritic membrane. However, in a similar current clamp experiment, the nonlinear behaviors measured from the soma are caused by the voltage responses of both the somatic and the dendritic membranes. Due to this asymmetry between voltage clamp and current clamp, there is no obvious way to predict the voltage clamp non- linear response from the current clamp nonlinear response nor the converse. It is well known in the linear case that the admittance from voltage clamp, Y (f ) = , I(f ) V (f ) , that is Y = Z−1 is the inverse of the impedance from current clamp Z (f ) = V (f ) I(f ) 7 V (f1+f2) I(f1) I(f2) where I or V refers to Fourier transforms of I and V, respectively. In the nonlin- ear case, with nonoverlapping frequencies, the quadratic response for the voltage clamp is defined by QSA as Bvc (f1, f2) = γf1,f2 and for the current clamp I(f1+f2) V (f1) V (f2) by Bcc (f1, f2) = γf1,f2 , where γf1,f2 is a symmetry factor (Magnani and Moore, 2010). In general Bvc (f1, f2) and Bcc (f1, f2) are not reciprocally equivalent because of the asymmetry discussed above. Hence, this is an important concep- tual difference between linear and nonlinear analysis, which also plays a role in interpretation of current versus voltage clamp experiments. The terms, V (f1 + f2) and I (f1 + f2) where f1 and f2 are either positive or neg- ative frequencies refer to all complex values of V and I at the interactive and harmonic quadratic frequencies. The quadratic response function can then be rep- resented as magnitude and phase plots versus f1 and f2. Matrix reduction Since Bvc (f1, f2) is difficult to interpret, it is convenient to reduce it to a diagonal matrix through eigendecomposition methods. Similar methods have been used in quantitative neuronal analyses, for example singular value decomposition (Lewis et al., 2002) or principal component analysis (Haas et al., 2007). To make this reduction, it is important to note that the matrix Qvc obtained by row flipping Qvc (f1, f2) = Bvc (−f1, f2) is actually Hermitian. As a consequence, Qvc can be reduced to a diagonal matrix D such that Qvc = P ∗DP where P is a unitary ma- trix and P ∗ its complex conjugate transpose. The unitary matrix P contains no information about the magnitude and can be viewed as a kind of generalization of the phase. The magnitude is entirely encoded in the diagonal matrix D. The ele- ments of D are called eigenvalues. Each column of the matrix P ∗ is a special vector called an eigenvector, whose coordinates are expressed relatively to the stimulus frequencies. The amplitude of each eigenvalue indicates the relative contribution of the corresponding eigenvector to the quadratic response. Practically, the eigen- decomposition can be interpreted as a reduction of the quadratic neuronal function to a set of quadratic filters in which eigenvalues play the role of amplitudes. The responses at the stimulus frequencies are shown in Figure 1A as interpolated black points. The color coded points illustrate the smaller amplitude current re- sponses at the quadratic frequencies. At low frequencies both the linear and non- linear responses are quite comparable despite the fact that the standard deviation of the imposed voltage clamp stimulus was only 2.85 mV. In Figure 1B, each cell (i, j) of the QSA matrix encodes the magnitude of the corresponding quadratic interaction (fi, fj). Informally, the QSA matrix can be viewed as a quadratic generalization of the admittance, thus it is an intrinsic characterization of the 8 measured neuronal response. The interpolations were performed by the MATLAB command GRIDDATA (linear method) in order to represent the responses in 3D color plots over a continuous range of frequencies. As a characteristic function, the QSA matrix is independent of the stimulus amplitude, however the current re- sponses become proportionally insignificant as the stimulus approaches zero. The responses of higher orders are insignificant for stimulation amplitudes below ±3 mV. The QSA matrix, Qvc, is the core mathematical object which encodes the total effect of the pairwise interactions at the interactive frequencies. Each matrix cell is located at the intersection of two input frequencies. In this way, the one dimensional admittance function Y (f ) is generalized to a two dimensional quadratic function Qvc (f1, f2). The QSA along with linear analysis can be used to reconstruct the signal using Y and Qvc. Figure 1B shows that the maximum amplitude for the interactive frequencies occurs at the intersection of 2 Hz and 10.4 Hz (f1 + f2 ). The eigendecomposition of these data strongly suggests that depolarized type D neurons are dominated by a single eigenvalue as illustrated in Figure 1C. Even when two eigenvalues are required to adequately describe the response, the eigendecom- position provides a remarkably compact representation of the nonlinear response that otherwise can only be quantitatively described by complex differential equa- tions involving large numbers of dendritic compartments. It would appear that the neuronal function becomes more complex, in the sense of information processing, as the number of significant eigenvalues increases. Matrix summation The previous matrix reduction has the great advantage of being reversible, such that from P and D one can exactly recover Qvc. There exists a coarser simplifi- cation by summing the QSA matrix by columns to obtain a vector indexed by the stimulation frequencies. The summation is defined as follows : (cid:88) R (j) = Qvc (i, j) i The values of R are illustrated in Figure 1D. The advantage of the R functions is that they can be presented as classical Bode plots. Moreover, each R function can be intuitively interpreted as a measure of the influence of each individual stimulation frequency on the nonlinear responses involving its interaction with all other stimulation frequencies. Hence, matrix summation is especially well suited to superimpose and compare with the piecewise quadratic analysis for different steady state responses. Figure 2A shows a reconstruction of the measured currents with the first and second order responses. The contribution of the nonlinear frequencies can be quantitatively 9 Fig. 1: Voltage clamp measurement at -50 mV on a type D neuron. (A) Amplitude of the evoked linear current responses (black points with interpolation line) and nonlinear frequency components (colored points). The stimulation show an increasing amplitude as the frequency increases. The nonlinear interactions are shown as individual points for fi + fj in red, fi − fj in blue, 2fi in green where fi and fj are positive. (B) Three dimensional plot of the magnitudes of the interpolated QSA matrix for a continuous range of frequencies. Each axis is indexed by 8 negative followed by 8 positive frequencies, which are shown as a 16x16 square matrix. (C) Eigenvalues of the QSA matrix, illustrating the dominance of one eigenvalue. The abscissa is shown with numbers indicating each of the 2N eigenvalues by decreasing magnitudes. (D) Summation R (j) showing the influence of each stimulation frequency in the quadratic responses. The voltage command was a multsinusoidal stimulation, which had a standard deviation of 2.85 mV. For all plots, the frequency components were computed with the MATLAB command FFT divided by the number of points. The current was measured in nA and the voltage in mV. The stimulation was constructed from the nonoverlapping set (0.2, 0.8, 2, 3.4, 5.8, 10.4, 13.4, 17.8) in Hz. 10 Fig. 2: Nonlinear and linear responses of the voltage clamped type D neuron of Figure 1 . (A) Superposition of the averaged data with the first (green) and second order (red) responses to describe the total neuronal response (blue). (B) Superposition of only first order responses for four traces. (C) Superposition of only second order responses for four traces which provides an indication of the synchrony and reproducibility of the second order components in the data for significantly large amplitudes. (D) Linear impedance interpolated from the stimulating frequencies. 11 expressed by comparing the ratios of the spectral energy of the linear versus linear + quadratic responses to the total spectral energy for the entire range of frequen- cies. Since the maximum stimulus frequency was 17.8 Hz, a frequency range of 36 Hz > 2 x 17.8 was selected for the total sum. Clearly, the quadratic reconstruction is much more accurate than the linear one, which is confirmed by evaluating the signal energy, 61% for the linear analysis against 96% for the quadratic analysis. The discrepancy between the second order ratio and 100% shows how well a second order approximation fits the total response, as well as an indication of the presence of other higher order responses for the frequency range selected. Higher order re- sponses clearly are more prominent with a depolarization due to the augmentation of the nonlinear behavior of both second and higher order responses. Since the second order responses have relatively low amplitudes it is essential to ensure that the contributions of other noise sources, such as synaptic events or membrane ion channel fluctuations, do not significantly contribute to the observed responses at the harmonic and interacting frequencies. The synchrony and repro- ducibility of the data is shown by superimposing first order (Figure 2B, delta1) and second order (Figure 2C, delta2) computed responses for four sequential measure- ments. A time invariance correlation function was used to determine an optimal stimulus amplitude, which is large enough to overcome the spontaneous noise of the neuron and not too great to evoke significant higher order responses (Magnani and Moore, 2010). Finally, the usual linear impedance, Z (f ) = , is shown in Figure 2D. V (f ) I(f ) Results Type D neurons It is difficult to get accurate measurements of the quadratic responses in current clamp for the type D neurons, mainly due to uncontrolled spontaneous oscillations. Thus, voltage clamp experiments were done to control the oscillations and measure the negative current associated with the persistent sodium conductance gNaP. The nonlinear responses evoked by 5-10 mV voltage clamp stimuli can also be blocked (not shown) by riluzole as described in an earlier steady state piecewise linear analysis of gNaP (Idoux et al., 2008). Figure 3A shows the increased linear impedance magnitude evoked by progressive depolarizations. Since gNaP can be seen as a negative conductance, its activation by depolarized membrane potentials reduces the total conductance of the cell, which leads to an increase of impedance. The nonlinear responses are indicated by eigenvalues (Figure 3B) and R summation (Figure 3C), whose magnitudes clearly 12 increase with depolarization. For this neuron, the spectral energy analysis led to ratios 99% and 100% at -60 mV, and 77% and 98% at -50 mV for the linear versus linear + quadratic responses, respectively. Type B neurons It has been shown previously that type B neurons have a prominent gNaP, which often leads to net inward current carried by Na+ for a limited range of voltage clamped depolarized membrane potentials (Idoux et al., 2008). Voltage clamped data in Figure 4 illustrate that type B nonlinear responses are significantly en- hanced by depolarization and in addition, often show a resonant enhancement of the impedance. The impedance shows a maximum at an intermediate depolariza- tion and a shift to a higher resonance frequency with further depolarization. In contrast, Figures 4B and C show that both the eigenvalues and the R summation values increase monotonically with depolarized membrane potentials. Interestingly, the number of significant eigenvalues required to describe the nonlinear response is generally two or more, unlike the single eigenvalue usually needed for type D neurons. For this neuron, the spectral energy analysis led to ratios 99% and 100% at -60 mV, and 82% and 96% at -41 mV for the linear versus linear + quadratic responses, respectively. Effect of gNaP and NMDA activation Nonlinear responses are likely to be enhanced by the activation of dendritic NMDA receptors, which would occur during synaptic activity of the neural integrator net- work. NMDA activation could trigger dendritic bistable responses that contributes to maintaining a particular firing rate after an input impulse. Figures 5B and C show that the addition of NMDA clearly enhances nonlinear response of a type D neuron consistent with the trigger hypothesis. The mean value of R and the maximum eigenvalue for all neuronal types increased significantly with the addi- tion of NMDA. TTX reduces the nonlinear effects of NMDA to control values or less, however it is still capable of inducing potential oscillations in current clamp (not shown). Control (with gNaP, without NMDA) versus NMDA + TTX treated neurons (without gNaP, with NMDA) did not show a significant p-value. Thus, NMDA and gNaP are synergistic in their action, where the combined effects are generally greater than either alone. Since the normal physiological activation of NMDA receptors is due to transient synaptic currents, the non-inactivation of gNaP and membrane potential bistability contribute to the maintenance of a depolarized potential. In conclusion gNaP appears to be essential for sustained nonlinear ef- fects induced by NMDA activation and thus would be critical for the operation of 13 Fig. 3: Voltage clamp experiments on a PHN type D neuron show the effect of an activated gNaP by comparing both the linear and quadratic voltage clamp responses at -70, -60, -55, -50 mV (blue, red, green, and black respectively). (A) Linear impedance, however it should be noted that the stimulus is a controlled voltage that leads to a current response. The depolarization markedly increases the impedance. (B) Eigenvalues of the QSA matrix. The abscissa is shown with numbers indicating each of the 2N eigenvalues by decreasing magnitudes. (C) Magnitude of R summations of the corresponding QSA matrix. They also increase during a depolarization. Statistically in five neurons the maximum eigenvalue increased for a change of membrane poten- tial of -60 to -50 mV from 0.002±0.001 to 0.010±0.001 with p-value=0.00004. The mean value of R±SD increased from 0.005±0.003 to 0.023±0.004 with p-value=0.0003. 14 Fig. 4: Voltage clamped type B neuron at -75, -60, -46, -41 mV (blue, red, green, black, respectively). (A) Linear impedance showing resonance induced by the depolarization. (B) Multiple eigenvalues increasing with depolarization in contrast to a single dominant eigen- value observed for type D neurons. The abscissa numerically labels the eigenvalues by decreasing magnitude. (C) Magnitude of R summations. These plots show that nonlinear responses are present at both hyperpolarized and depolarized membrane potentials. Statistically in six neurons, for a positive membrane potential change of 5 mV in the range -60 to -40 mV, the maximum value of the eigenvalue increased from 0.004±0.003 to 0.014±0.009 with p-value=0.008 and the mean value of R±SD increased from 0.009±0.006 to 0.03±0.02 with p-value=0.01. 15 the neural integrator. Type D and B neuronal model simulations In order to provide an interpretation of these experimental results, numerical sim- ulations have been done using previously published models of types D and B for rat PHN neurons (Idoux et al., 2008). These models have been constructed from a piecewise linear analysis to fit parameters of nonlinear differential equations in voltage clamp. Unless otherwise indicated, both models had three uniformly dis- tributed voltage dependent conductances, gK, gNaP and gH with a soma and eight dendritic compartments. Typically the gH of type D neurons is quite small and could be neglected. Simulations done with the published average parameter values were consistent with the voltage clamp data, however some appropriate modifications of the average parameters were done for the comparison with the data for the individual neurons of Figures 3 and 4. Four potentials are shown to cover the range observed in the experiments. The type D data and model generally show a dominant eigenvalue (see -50 mV in Table 1 and Figure 6). The simulations of the type B model in Figure 7 show behavior similar to the data of Figure 4, showing two or more significant eigenvalues compared to an essentially single dominant eigenvalue for the type D model (Figure 6). However Tables 1 and 2 indicate that the number of significant eigenvalues for both neuronal types is dependent on both the membrane potential and specific parameter values. For example, at -40 mV with or without NMDA, Figure 5 B shows a type D neuron with multiple significant eigenvalues (also, see Table 1). In addition, the monotonic increase in the R summation of the type B model non- linear responses with depolarization contrasts with the peaking of linear impedance increase similar to that observed in the type B neuron of Figure 4. In general the impedance increase with depolarization is caused by the activation of the gNaP neg- ative conductance that balances the other positive conductances. An impedance maximum, often with a resonance, occurs because the impedance decreases with further depolarization due to an increased gK. The nonlinear responses of type B model simulations, in contrast to type D, show that the quadratic response is greatly reduced in magnitude at -60 mV, in part because the type B model has a greater density of voltage dependent gNaP con- ductances on a relatively more compact dendritic structure than found for type D neurons. In addition, the nonlinear responses in the type B model, as in the data of Figure 4, are enhanced at hyperpolarized potentials directly due to the gH conductance. The use of nonoverlapping frequencies is required for the construction of QSA 16 Fig. 5: Effect of NMDA (red) and NMDA+TTX (green) compared to control (blue) on a voltage clamped type D neuron at -40 mV. (A) Effect of 25 µM NMDA on linear impedance. The impedance decreases with NMDA (red) and increases with 2 µM TTX (green) compared to the control (blue) (B) The NMDA plots (red) illustrate that the nonlinear responses are increased by the presence of NMDA. Moreover, TTX (green), which blocks gNaP, when added to NMDA shows decreased eigenvalues. The abscissa numerically labels the eigenvalues by decreasing magnitude. (C) Effect of NMDA and NMDA+TTX on R summation. The R values are in inverse order to the impedance magnitudes. Statistically, pooled data from five neurons (three of type B and two of type D) in the presence versus absence of 25-50 µM NMDA over the range -66 to -40 mV, showed a significant difference in the mean values of R summation (p-value = 0.026). Similarly the maximum eigenvalues were significantly different with a p-value of 0.035. TTX reduces the nonlinear effects of NMDA to control values or less (p-value = 0.09). An insignificant p-value of 0.2 was found comparing control with NMDA + TTX treated neurons. 17 Fig. 6: Model of voltage clamped type D PHN neuron. Simulations were done at -70, -60, -55, -50 mV (blue, red, green, black, respectively). (A) Linear impedance. The depolarization markedly increases the linear impedance. (B) Eigenvalues of the QSA matrix. The abscissa numerically labels the eigenvalues by decreas- ing magnitude. (C) Magnitude of R summations of the corresponding QSA matrix. The following parameter values were used: csoma = 20.5 pF; aratio = 3.77; elength = 0.54; gleak = 1.2 nS(1.37 nS); vleak = -53 mV; gk = 1.18 nS; vn = -32 mV; sn = 0.06(mV)−1 (0.05 (mV)−1); tn = 10 msec; vk = -87 mV; gH = 0; sm = 0.06 (mV)−1; tm = 150 µsec; vNa = 77 mV; vm = -38 mV (-35 mV); gNaP= 0.9 nS (0.6 nS). Parameter values shown in parentheses are default values defined in Idoux et al., 2008. 18 Fig. 7: Model of voltage clamped type B neuron. Simulations were done at -76, -60, -46, -41 mV (blue, red, green, black, respectively). (A) Linear impedance. The magnitude shows a resonant maximum at -46 mV. (B) Eigenvalues of the QSA matrix. The abscissa numerically labels the eigenvalues by decreas- ing magnitude. (C) Magnitude of R summations of the corresponding QSA matrix. The parameter values were as follows: csoma = 0.0000995 µF (0.0000265); aratio = 2.6 (2.85); elength = 0.31 (0.37); gleak = 0.001297 µS (0.000878); vleak = -56.61 µS (-54.9); gk = 0.005555 µS (0.0024); vn = -38.7 mV (-35); sn = 0.07715/mV (0.05); tn = 0.66 sec (.09); vk = -87 mV; gH = 0.00644 µS (0.00591); vq = -79.7 mV (-63.9); sq = -0.0667/mV (-0.0647); tq = 1.75 sec (8.43); vk2 = -43 mV; gNaP = 0.00147 µS (0.00131); vm = -35 mV; sm = 0.05/mV (.056); tm = 0.00015 sec; vnmda = 77 mV; Parameter values shown in parentheses are default values from Idoux et al., 2008. 19 soma-Na dendrite-Na Vm Rmean µ1 µ2 µ2/µ1(%) 1 1 0 0.01 1 1 1 0 1 0.01 -40 -50 -50 -50 -50 0.0162 +0.0160 0.0095 0.0011 +0.0011 0.0075 0.0012 -0.0049 -0.0095 +0.0015 -0.0002 -0.0074 +0.0012 -0.0012 +0.0003 31 16 18 16 25 Tab. 1: Type D model simulations. The voltage clamped somatic membrane potential is given along with presence or absence of conductances indicated by 1 or 0. The value of gNaP in either the soma or all dendritic compartments was reduced to 1% indicated by 0.01 in the labeled columns. The mean values of R summation and the two largest eigenvalues µ1 and µ2 are given for each simulation. The last column is µ2 represented as the percentage of the absolute µ1 value. A positive µ1 occurs when gK has a greater effect than gNaP. matrices from experimental measurements. Nevertheless, it is possible to do a coarse interpolation for other frequencies as plotted in Figure 8. The interpolated QSA matrices for type D and B neurons show striking differences, which also occur in their mathematical models. Figures 8A and 1B show local peaks in the QSA plots at two membrane potentials for the same type D neuron, which are comparable to the type D model simulation in Figure 8B. By contrast the type B neuron and model (Figure 8C, D) show a prominent low frequency peak distinctly different than observed in type D neurons. The QSA matrix clearly provides significantly more information about the neuronal behavior than linear parameters and has the possiblilty of being extended to all frequencies of interest. In order to test the hypothesis that the dendritic conductances are mainly respon- sible for the nonlinear responses, simulations were done with gNaP reduced in the soma and unchanged in the dendrite or the converse as indicated in Tables 1 and 2. In type D neurons, decreasing the soma gNaP has a lesser effect (79%) on the R summation compared with decreasing the dendritic gNaP (13%). Simula- tions of type B model neurons showed a lesser effect for a gNap reduction in the more compact dendrite (34%). Such simulations support the hypothesis that the nonlinear responses measured under voltage clamp conditions are dominated by dendritic responses, while the linear responses are determined by both the soma and dendrite. Since the somatic region has good voltage space clamp control, one would expect a dominant linear current response from the somatic conductances at all membrane potentials, however the significantly larger dendritic membrane area is not space clamped leading to uncontrolled voltage excursions and greater nonlinearites. In conclusion, these simulations suggest that nonlinear responses in voltage clamped neurons are dominated by active dendritic structures, when their electrotonic lengths are above 0.3. soma-Na dendrite-Na Vm Rmean 0.0140 0.00652 0.00104 +0.0018 0.00467 0.00221 -0.0141 +0.0027 -0.0063 +0.0018 -0.0008 -0.0044 +0.0020 -0.0019 +0.0015 1 1 0 0.01 1 µ1 µ2 1 1 0 1 0.01 -40 -50 -50 -50 -50 20 µ2/µ1(%) 19 29 44 45 79 Tab. 2: Type B model simulations as in Table 1. The mean values of R summation and the two largest eigenvalues µ1 and µ2 are given for each simulation. When gNaP in the dendrite is reduced, the negative µ1 decreases and becomes as large as the unchanged positive µ2 (79%), which suggests structurally different quadratic filters for type B compared to type D models. Fig. 8: Interpolated QSA plots of type D and B neurons. (A) type D neuron of Figure 3 at -50 mV. (B) Type D model of Figure 6 at -50 mV. (C) Type B neuron of Figure 4 at -46 mV. (D) Type B model of Figure 7 at -46 mV. 21 Discussion The neurons of the prepositus hypoglossi nucleus (PHN) provide a useful system to investigate nonlinear behavior, such as persistent activity to maintain eye position. The oscillatory character of some of these neurons is similar to that observed in the stellate neurons in layer IV of entorhinal cortex (Haas and White, 2002; Schreiber et al., 2004), which are also involved in the processing of orientational information. PHN neurons are part of the brainstem that receives head velocity signals and integrates them to control eye position for the stabilization of an image at the center of the visual field during head rotation. This specific processing is called neural integration (Aksay et al., 2007) due to an analogy with integration in mathematical calculus. In order to understand how the PHN neural network can perform neural integra- tion, it is important to understand the biophysical properties of individual neurons involved in the circuitry. Single neurons of the PHN show oscillatory and bistable nonlinear properties that are likely to be involved in the operation of the neural integrator. Since models based on recurrent excitation, even including lateral in- hibition, are not sufficiently robust, a number of theoretical papers have suggested that the nonlinear properties of individual neurons are essential for the network be- havior of the neural integrator (Koulakov et al., 2002; Goldman et al., 2003). Thus, the finding that oscillatory nonlinear behavior is clearly present in the neurons in- volved in the eye movement circuitry (Idoux et al., 2006) lends strong support to these theoretical notions. Dendritic characteristics (Johnston and Narayanan, 2008) are potentially critical for the function of these neurons, which in addition have also shown to be different for the two main classes of PHN neurons (Idoux et al., 2008). Thus, type D oscillatory neurons have bistability properties, which are consistent with neural integrator models that rely on remote dendritic processing (Goldman et al., 2003; Idoux et al., 2008). The experiments and analysis in this paper strongly support the hypothesis that type D neurons have persistent sodium channels in their dendrites, which would promote remote bistable potentials shifts leading to persistent activity. Type B neurons have an even higher density of persistent sodium channels and their passive electrotonic length is less than type D. As a consequence, the type B neurons at moderate depolarizations would have a more uniform potential throughout the less isolated dendritic tree. The gNaP conductance in the dendrites can easily be activated by the synaptic stimulation of the NMDA receptors likely to occur during normal physiological activity. In addition, NMDA activation enhances total nonlinearities, which are shown for a type D neuron and also observed in type B neurons (see Figure 5). In current clamp conditions, both types of neurons show marked potential oscillations 22 in the presence of 25-50 µM NMDA (Idoux et al., 2006). Clearly the level of NMDA activation expected during synaptic activation would be much less, however it is likely to be sufficient to evoke significant nonlinear responses due to the gNaP in both type B and D neurons. In this regard, the neural integrator model of Koulakov et al. (2002), depends on NMDA synapses that activate gNaP dependent bistable states in dendritic compartments. The quantitative measurement of the biophysical properties of intact neurons is seriously compromised by the inherent inability to voltage clamp the electrotonic structure of the dendritic tree. In general whole cell measurements are restricted to patch clamp electrodes placed in the soma, which makes it difficult to infer the remote properties of the dendrites. Previous piecewise linear analyses have permitted the development of realistic neuronal models, however it has been diffi- cult to separate the properties of the dendrites from the soma unless patch clamp electrodes can be placed in the dendrites. The new approach described in this pa- per takes advantage of the space clamp problem associated with voltage clamping neurons. The quadratic analysis has shown that it is possible to characterize the nonlinear behavior of the uncontrolled dendritic membrane voltage responses while maintaining voltage clamp control of the somatic membrane. When the dendritic electrotonic structure is remote with relatively large surface areas compared to the soma, the nonlinear behavior of the dendritic membrane dominates the soma. This is especially the case in a voltage clamp experiment because of the lack of dendritic potential control. It was found that responses in the range of ±(5-10) mV could be well described by quadratic nonlinearities suggesting that nonlinearities of higher degrees only add marginal improvement. Thus, the quadratic response is likely to sufficiently capture most of the nonlinear behavior of neuronal systems except for extremely large synaptic inputs. The quadratic functions are quite sensitive to the mean membrane potential and appear to be valid for a range of sinusoidal inputs. This behavior extends significantly the validity of the quantitative quadratic description. The quadratic functions are computed on particular sets of nonoverlapping frequen- cies, which can be interpolated over a continuous range of frequencies as illustrated in Figures 1B and 8. Thus, they provide a significantly concise description of the neuronal behavior and could potentially be used as computational devices that would be independent of nonlinear differential equations. Practically, this could be an alternative approach to large scale neural network simulations. The estimation of the parameters of both the voltage dependent conductances and the electrotonic structure have shown quantitative differences between type B and D neurons (Idoux et al., 2008). In addition, the nonlinear analysis in this paper suggests that the number of significant eigenvalues is greater for the type B versus the type D neurons and their individual models. Thus, the measured nonlinear- 23 ities seems to be structurally different between type B and D, namely the two corresponding types of quadratic functions are intrinsically different (Figure 8). In general the dominant eigenvalue was negative, which is related to the negative slope conductance due to gNaP. Type D model simulations at large depolarizations (see Table 1 at -40 mV) show that the maximum eigenvalue (µ1) is positive, consistent with a positive slope conductance due to an increased outward potassium current. In contrast, the greater gNaP of the type B model maintains a negative µ1 at -40 mV (Table 2). Both type D and B models show positive µ1 values if gNaP is totally removed from the soma and dendrite. In conclusion, the work described in this paper provides a novel way to concisely quantify the fundamental nonlinearities underlying of individual neurons. By allow- ing rigorous comparison of any neuronal model with the behavior of real neurons, it makes possible to show that nonlinear responses in voltage clamp are dominated by the active dendritic structure. A determination of the molecular basis of the eigenvalue analysis should provide a better understanding of how neurons use their remarkable nonlinear properties in information processing. References Aksay E, Olasagasti I, Mensh BD, Baker R, Goldman MS, Tank DW (2007) Func- tional dissection of circuitry in a neural integrator. Nat Neurosci 10:494 -- 504. Boyd S, Tang Y, Chua L (1983) Measuring volterra kernels. IEEE Transactions on Circuits and Systems 30:571 -- 577. Erchova I, Kreck G, Heinemann U, Herz AVM (2004) Dynamics of rat entorhinal cortex layer ii and iii cells: characteristics of membrane potential resonance at rest predict oscillation properties near threshold. J Physiol 560:89 -- 110. Fishman HM, Poussart DJ, Moore LE, Siebenga E (1977) K+ conduction descrip- tion from the low frequency impedance and admittance of squid axon. J Membr Biol 32:255 -- 290. FitzHugh R (1983) Sinusoidal voltage clamp of the hodgkin-huxley model. Biophys J 42:11 -- 16. French AS (1976) Practical nonlinear system analysis by wiener kernel estimation in the frequency domain. Biological Cybernetics 24:111 -- 119. Goldman MS, Levine JH, Major G, Tank DW, Seung HS (2003) Robust persis- tent neural activity in a model integrator with multiple hysteretic dendrites per neuron. Cereb Cortex 13:1185 -- 1195. 24 Gutkin BS, Ermentrout GB (1998) Dynamics of membrane excitability determine interspike interval variability: a link between spike generation mechanisms and cortical spike train statistics. Neural Comput 10:1047 -- 1065. Haas JS, Dorval AD, White JA (2007) Contributions of ih to feature selectivity in layer ii stellate cells of the entorhinal cortex. J Comput Neurosci 22:161 -- 171. Haas JS, White JA (2002) Frequency selectivity of layer ii stellate cells in the medial entorhinal cortex. J Neurophysiol 88:2422 -- 2429. Hodgkin AL, Huxley AF (1952) A quantitative description of membrane current and its application to conduction and excitation in nerve. J Physiol 117:500 -- 544. Idoux E, Eugene D, Chambaz A, Magnani C, White JA, Moore LE (2008) Con- trol of neuronal persistent activity by voltage-dependent dendritic properties. J Neurophysiol 100:1278 -- 1286. Idoux E, Serafin M, Fort P, Vidal PP, Beraneck M, Vibert N, Muehlethaler M, Moore L (2006) Oscillatory and intrinsic membrane properties of guinea pig nucleus prepositus hypoglossi neurons in vitro. J Neurophysiol 96:175 -- 196. Izhikevich EM (2002) Resonance and selective communication via bursts in neurons having subthreshold oscillations. Biosystems 67:95 -- 102. Johnston D, Narayanan R (2008) Active dendrites: colorful wings of the mysterious butterflies. Trends Neurosci 31:309 -- 316. Koulakov AA, Raghavachari S, Kepecs A, Lisman JE (2002) Model for a robust neural integrator. Nat Neurosci 5:775 -- 782. Lewis ER, Henry KR, Yamada WM (2002) Tuning and timing of excitation and inhibition in primary auditory nerve fibers. Hear Res 171:13 -- 31. Magnani C, Moore LE (2010) Quadratic sinusoidal analysis of voltage clamped neurons. arXiv:1007.0858v1. Marmarelis PZ, Naka KI (1973) Nonlinear analysis and synthesis of receptive-field responses in the catfish retina. i. horizontal cell leads to ganglion cell chain. J Neurophysiol 36:605 -- 618. Moore LE, Chub N, Tabak J, O'Donovan M (1999) Nmda-induced dendritic J Neu- oscillations during a soma voltage clamp of chick spinal neurons. rosci 19:8271 -- 8280. 25 Moore LE, Fishman HM, Poussart DJ (1980) Small-signal analysis of k+ conduc- tion in squid axons. J Membr Biol 54:157 -- 164. Murphey CR, Moore LE, Buchanan JT (1995) Quantitative analysis of electrotonic structure and membrane properties of nmda-activated lamprey spinal neurons. Neural Comput 7:486 -- 506. Schetzen M (2006) The Volterra and Wiener Theories of Nonlinear Systems Krieger Publishing Company. Schreiber S, Erchova I, Heinemann U, Herz AVM (2004) Subthreshold resonance explains the frequency-dependent integration of periodic as well as random stim- uli in the entorhinal cortex. J Neurophysiol 92:408 -- 415. Vervaeke K, Hu H, Graham LJ, Storm JF (2006) Contrasting effects of the persis- tent na+ current on neuronal excitability and spike timing. Neuron 49:257 -- 270. Victor J, Shapley R (1980) A method of nonlinear analysis in the frequency domain. Biophys J 29:459 -- 483.
1701.01311
3
1701
2017-01-09T14:21:38
Extracting the Groupwise Core Structural Connectivity Network: Bridging Statistical and Graph-Theoretical Approaches
[ "q-bio.NC", "cs.DM" ]
Finding the common structural brain connectivity network for a given population is an open problem, crucial for current neuro-science. Recent evidence suggests there's a tightly connected network shared between humans. Obtaining this network will, among many advantages , allow us to focus cognitive and clinical analyses on common connections, thus increasing their statistical power. In turn, knowledge about the common network will facilitate novel analyses to understand the structure-function relationship in the brain. In this work, we present a new algorithm for computing the core structural connectivity network of a subject sample combining graph theory and statistics. Our algorithm works in accordance with novel evidence on brain topology. We analyze the problem theoretically and prove its complexity. Using 309 subjects, we show its advantages when used as a feature selection for connectivity analysis on populations, outperforming the current approaches.
q-bio.NC
q-bio
Extracting the Groupwise Core Structural Connectivity Network: Bridging Statistical and Graph-Theoretical Approaches Nahuel Lascano1,2, Guillermo Gallardo1, Rachid Deriche1, Dorian Mazauric3, and Demian Wassermann1 1 Athena EPI, Universit´e Cote d'Azur, Inria, France 2 Computer Science Department, FCEyN, Universidad de Buenos Aires, Argentina 3 ABS EPI, Universit´e Cote d'Azur, Inria, France Abstract. Finding the common structural brain connectivity network for a given population is an open problem, crucial for current neuro- science. Recent evidence suggests there's a tightly connected network shared between humans. Obtaining this network will, among many ad- vantages, allow us to focus cognitive and clinical analyses on common connections, thus increasing their statistical power. In turn, knowledge about the common network will facilitate novel analyses to understand the structure-function relationship in the brain. In this work, we present a new algorithm for computing the core struc- tural connectivity network of a subject sample combining graph theory and statistics. Our algorithm works in accordance with novel evidence on brain topology. We analyze the problem theoretically and prove its complexity. Using 309 subjects, we show its advantages when used as a feature selection for connectivity analysis on populations, outperforming the current approaches. Keywords: Group-wise connectome, core graph problem, brain connec- tivity, diffusion MRI 1 Introduction Isolating the common brain connectivity network from a population is a main problem in current neuroscience [3,7,10]. Recent evidence suggests that there's a common and densely connected brain connectome across humans [2]. In this work we present a new approach for selecting these common connections, combining recent topological hypotheses [2] and current methods [7,10]. Finding the common brain connectome across subjects has the potential to increase our understanding of the relationship between function and struc- ture in the brain. This relationship is one of the main open questions in neu- roscience [3,5]. Moreover, knowledge about the most common connections in a population will facilitate clinical and cognitive Diffusion MRI analyses by re- ducing the number of surveyed connections, increasing the statistical power of those analyses. Finding the common connectome will also allow us to increase 7 1 0 2 n a J 9 ] . C N o i b - q [ 3 v 1 1 3 1 0 . 1 0 7 1 : v i X r a 2 N. Lascano et al. our knowledge about the brain structure by comparing core networks across different populations. We formalize the problem of selecting the common connections combining graph theory and statistics. Then, we prove that the problem is NP-Hard and propose a polynomial-time algorithm to find approximate solutions. To do this, we develop an exact polynomial-time algorithm for a relaxed version of the problem and prove the algorithm's correctness and complexity. Currently, the most used algorithm to extract a population's core structural connectivity network (CSNC) [7] uses an statistical approach: first, compute a connectivity matrix for each subject; then, analize each connection separately with a hypothesis test, using as null hypothesis that that edge is not present in the population; finally, construct a binary graph with the edges for which the null hypothesis was rejected. The main problem of Gong et al.'s [7] algorithm is that the resulting graph can be a set of disconnected subgraphs. Moreover, recent studies have shown that the brain has a core network tightly connected and a sparsely connected outer one [2]. In other words, this approach ignores the resulting network's topology. Performing statistical analyses in a feature set chosen by hypothesis testing incurs in the double dipping problem [8]. A newer approach to solve the CSNC problem, designed by Wassermann et al. [10], uses graph theory to get a connected CSCN: first, compute a binary connectivity graph for each subject using a threshold; for each possible connec- tion compute the "cost" of including or excluding it from the common graph by evaluating in how many subjects that connection is present; finally, construct the binary graph with all the edges that is "cheaper" to include than to ex- clude and connect the resulting graph if it's disconnected, using the minimum possible cost. This algorithm guarantees that the resulting graph is connected, but the connection binarization discards significant information for the resulting common network. In other words, it discards information of the probability of each connection being in the brain. This is problematic because the resulting graph may include edges for which tractography assigned a very low existence probability across subjects. Also, the outer part of the brain, the connections which do not result in the core network, should also be sparsely connected [2], which this algorithm does not enforce. In this work we propose, for the first time, a polynomial-time algorithm to ob- tain the CSCN of a population addressing the issues listed above. Our algorithm combines the recent graph-theoretical approach [10] with the statistical aware- ness of the most popular one [7]. We start by formalizing the problem, which allow us to prove that it's NP-Hard. Then, we propose a first algorithm that solves a relaxed version of the problem in an exact way, giving the best possible core graph for our formalization. Then, we adapt it to guarantee a connected result, agreeing with recent evidence on structural connectivity network topol- ogy [2, e.g.]. Finally, we validate our approach using 300 subjects from the HCP database and comparing the performance of the networks obtained by our new approach, Wassermann et al.'s [10] and Gong et al.'s [7] predicting connectivity values from handedness in the core network. Extracting the Groupwise Core Structural Connectivity Network 3 2 Definitions, Problems and Contributions We want to develop a new algorithm to extract the core structural connectivity network, a problem that implies working with different brains. Thus, the first thing we need to do is unify them into a common connectivity model. This allows us to model all brains with graphs in which each node represents a cortical or sub- cortical region, and each edge represents a white matter connection between two regions. We choose the Desikan parcellation [4] to uniformize the brain cortical and sub-cortical regions across subjects. To compute the connectivity matrices we use a probabilistic tractography al- gorithm, which outputs one matrix per subject. The resulting matrices represent the existence probability of a connection across parcels in each subject [5]. As these are symmetric, we interpret the matrices as weighted undirected graphs, sharing the node set across subjects. Formally, we represent a sample of N brain structural networks by N com- plete weighted graphs G1 = (V, E, w1), . . . , GN = (V, E, wN ) with a common node set V . We call G1, . . . , GN the sample graphs. Each graph Gi corresponds to a subject. Each vertex v ∈ V represents a cortical or sub-cortical region. Each edge e ∈ E = V × V represents a white matter bundle connecting two regions. Finally, the weight wi(e) is the connection probability for the edge e in the subject i obtained through tractography: w1(e), w2(e), . . . , wN (e) ∈ [0, 1] ∀e ∈ E. (1) Note that all graphs have the same ordered node set and all of them are com- plete: an edge weight, or connection probability, wi(e) of 0 represents an absent connection. Using this formalization we express the general core structural con- nectivity network problem as follows: find a core graph G∗ = (V ∗, E∗) densely connected such that G∗ keeps the more relevant connections E∗ ⊆ E in the sample and discards the less relevant ones, for some definition of relevance and density. For simplicity, once we select E∗ we can define V ∗ as V ∗ = {v ∈ V : ∃u ∈ V, (u, v) ∈ E∗} , (2) the set of nodes that the edges in E∗ cover. Then, we can reduce the problem of finding G∗ to find E∗ alone. We want a formalization of relevance that represents the probability that a connection is present across subjects. Thus, we choose to model the group- wise relevance, w∗(e) as the mean existence probability across subjects, factored by the standard deviation of these probabilities. In other words, w∗(e) is the number of standard deviations that the mean existing probability of connection e is larger than 0. We use w∗(e) as a statistical measure of edge presence across the population. Formally, where w(e) (cid:44) N(cid:88) wi(e) N , s(e) (cid:44) i=1 w∗(e) (cid:44) w(e) s(e) (cid:16) (cid:118)(cid:117)(cid:117)(cid:117)(cid:116) N(cid:88) i=1 (cid:17)2 w(e) − wi(e) N . (3) (cid:80) e∈E∗ w∗(e) E∗ (cid:80) e∈E\E∗ w∗(e) E∗ . . (4) (5) α(w∗, E∗) (cid:44) β(w∗, E∗) (cid:44) 4 N. Lascano et al. Note that w∗(e) is the statistic of a hypothesis z-test which assumes a media of 0 for the population weight of e. We choose the z-statistic because of the normal distribution's properties, e.g. linearity, even if other distributions, such as Beta distribution, may be more appropriate for modeling the probability. In any case, note that for the purpose of our contribution w∗ can be any function E → R which grows with the relevance of the edges in the sample. We also want a formalization that represents the density of the core subgraph. We use the relationship between the number of edges and the total statistical relevance w∗ that those edges sum: As we want also a sparse outer subgraph, we also define its density: Now we can express our objective informally as: choose E∗ such that α(w∗, E∗) (Eq. 4) is large and β(w∗, E∗) (Eq. 5) small. In accordance to recent evidence on the core network, we also want G∗ to be connected. Here, connected means that for every pair of vertices u, v in V ∗ there is a path of edges in E∗ from u to v. Let E c be the family of sets of edges that induce a connected graph. We now formalize the problem of finding this common graph G∗ in two different ways. • The optimization version consists in computing: f (w∗, E∗) = λα(w∗, E∗) − (1 − λ)β(w∗, E∗) (6) max E∗∈E c The parameter λ (between 0 and 1) can be adjusted to weight the density of the inner and the outer network. Note that if λ = 1, the solution to (6) only considers the density of the core network, and if λ = 0, it only considers the edges excluded of the core network. • Given A and B, the decision version consists in finding E∗ ⊆ E c such that: α(w∗, E∗) ≥ A β(w∗, E∗) ≤ B (7) Having formalized the Core Structural Connectivity Network into an opti- mization and a decision problem, we proceed with one of our main theoretical contributions: proving that the problem is NP-Complete. 2.1 CSCN Problem's NP-Completeness We have formalized the problem of the Core Structural Connectivity Network taking into account the density and connectedness of the core subgraph and the sparsity of the outer one. We will now prove that, with this formalization, the problem is NP-Complete. Extracting the Groupwise Core Structural Connectivity Network 5 Definition 1 (Core Structural Connectivity Network problem). Given G1 = (V, E, w1), G2 = (V, E, w2), . . . , GN = (V, E, wN ) weighted graphs (the sample graphs) with a common node set, a complete edges set (E = V × V ) and w1(e), w2(e), . . . , wN (e) ∈ R≥0 ∀e ∈ E weights of their edges, and given A, B real numbers, find G∗ = (V ∗, E∗) connected graph (the core graph) such that α(w∗, E∗) ≥ A β(w∗, E∗) ≤ B for α and β as defined in Eq. (4) and Eq .(5). Here we prove that the Core Structural Connectivity Network problem, called CSCN problem, is NP-complete. In our reduction, we use the Steiner Tree prob- lem [6], called ST problem in the following. Given an edge-weighted graph G(cid:48) = (V (cid:48), E(cid:48), w), a subset S ⊆ V (cid:48) of nodes, and a real k ≥ 0, ST problem consists (cid:80) in determining if there exists a connected subgraph H such that S ⊆ V (H) and e∈E(H) w(e) ≤ k. The decision version of ST problem is NP-complete even if Instance of ST problem. Consider any edge-weighted graph G(cid:48) = (V (cid:48), E(cid:48), w) 2 for every e ∈ E(cid:48). Given k ≥ 0, ST problem consists in de- such that w(e) = 1 termining if there exists a connected subgraph H such that S ⊆ V (H) and E(S) ≤ 2k. Without loss of generality, we assume that E(cid:48) ≥ 2k and that G(cid:48) is connected. all weights are equal [6]. Reduction. We construct the instance of CSCN problem as follows. Let s = S and let t ≥ 1 be any positive integer. Let G = (V, E, w∗) defined as follows. Let V = V (cid:48) ∪ {vi,j 1 ≤ i ≤ s, 1 ≤ j ≤ t} and E = V × V . Let S = {u1, . . . , us}. For every i, j, 1 ≤ i ≤ s, 1 ≤ j ≤ t, w∗ vi,j ,u = 0 for every u ∈ V \ {ui}. Furthermore, for every e ∈ E(cid:48), set w∗(e) = w(e) = 1 2 , and for every u, u(cid:48) ∈ V (cid:48) such that {u, u(cid:48)} /∈ E(cid:48), then set w∗ e = 0. Finally, we set A = s.t+k Lemma 1. If E∗ < s.t + 2k, then any solution for CSCN problem is not admissible because β(w∗, E∗) > B. Proof. Suppose that E∗ < s.t + 2k. In order to minimize(cid:80) s.t+2k and B = = 1 and w∗ 2 (E(cid:48)−2k) vi,j ,ui s.t+2k . 1 e∈E\E∗ w∗(e), E∗ must contain {{vi,j, ui} 1 ≤ i ≤ s, 1 ≤ j ≤ t} if E∗ ≥ s.t. (Otherwise we select a subset of this set of edges.) Indeed, by construction of G, we have w∗ = 1 for every i, j, 1 ≤ i ≤ s, 1 ≤ j ≤ t. Then, if E∗ − s.t > 0, E∗ must contain E∗ − s.t edges of E(cid:48), that is edges of E of weight 1 2 each. Recall that there are exactly s.t edges of weight 1, and the other edges have weight 0 or 1 2 . e∈E\E∗ w∗(e) = There are two cases. First, suppose that E∗ ≥ s.t. We get that(cid:80) 2 (E(cid:48) − (E∗ − s.t)). Since E∗ − s.t < 2k, then we get that(cid:80) 2 (E(cid:48) − (E∗ − s.t)) > 1 2 (E(cid:48)−(E∗−s.t)) (cid:80) e∈E\E∗ w∗(e) = 2 (E(cid:48) − 2k). Furthermore, since E∗ < s.t + 2k, . Thus, we proved that β(w∗, E∗) = 2 (E(cid:48)−2k) E∗ vi,j ,ui s.t+2k > 1 1 1 1 we get that e∈E\E∗ w∗(e) E∗ > B. 6 N. Lascano et al. Second, suppose that E∗ < s.t. We get that(cid:80) e∈E\E∗ w∗(e) = s.t − E∗ + E(cid:48) 2 (E(cid:48) − (E∗ − s.t)), we obtain the result by the 2 . Since s.t − E∗ + Finally, if E∗ < s.t + 2k, then there is no admissible solution for CSCN (cid:3) arguments described for the first case. E(cid:48) 2 > 1 problem. Lemma 2. If E∗ > s.t + 2k, then any solution for CSCN problem is not admissible because α(w∗, E∗) < A. Proof. Suppose that E∗ > s.t + 2k. In order to maximize (cid:80) e∈E∗ w∗(e), E∗ must contain {{vi,j, ui} 1 ≤ i ≤ s, 1 ≤ j ≤ t} and E∗ − s.t edges of E(cid:48). = 1 for every i, j, 1 ≤ i ≤ s, Indeed, by construction of G, we have w∗ 2 for every e ∈ E(cid:48). Recall that there are 1 ≤ j ≤ t. Furthermore, w∗(e) = 1 exactly s.t edges of weight 1, and the other edges have weight 0 or 1 2 . We get that α(w∗, E∗) = s.t+ 1 s.t+2k = A. Indeed, the average weight is lower when there are more edges of weight 1 2 (the number of edges of weight 1 is the same in both ratios). Finally, if E∗ > s.t + 2k, then there is no admissible solution for CSCN (cid:3) 2 (E∗−s.t) E∗ < s.t+k vi,j ,ui problem. By Lemma 1 and Lemma 2, we get the following corollary. Corollary 1. Any solution for CSCN problem is such that E∗ = s.t + 2k. We prove in Lemma 3 and in Lemma 4 that there is an admissible solution for CSCN problem if and only if there is an admissible solution for ST problem. Lemma 3. If there is an admissible solution for ST problem, then there is an admissible solution for CSCN problem. e∈E(H) w(e) = 1 graph such that S ⊆ V (H) and(cid:80) Proof. Suppose there is an admissible solution for ST problem. We prove that there is an admissible solution for CSCN problem. Let H be a connected sub- 2E(H) ≤ k. If E(H) = 2k, then set E∗ = E(H) ∪ {{vi,j, ui} 1 ≤ i ≤ s, 1 ≤ j ≤ t}. If E(H) < 2k, then set E∗ = E(H) ∪ {{vi,j, ui} 1 ≤ i ≤ s, 1 ≤ j ≤ t} ∪ F , where F ⊆ E(cid:48) such that F (cid:54)= E(H) = ∅, w∗(e) = 1 2 for every e ∈ F , and such that the graph induced by E(H) ∪ F is connected. The last condition comes from Corollary 1 in order to get the right number of edges in E∗. This condition is always possible to satisfy because G(cid:48) is connected. The graph induced by E∗ is connected. Indeed, H is an admissible solution for ST problem, E(H)∪ F is connected by construction, and {{vi,j, ui} 1 ≤ j ≤ t} is a set of edges all adjacent to ui ∈ S for every i, 1 ≤ i ≤ s. Furthermore, we get α(w∗, E∗) = (cid:80) e∈E∗ w∗(e) E∗ = s.t + k s.t + 2k = A Extracting the Groupwise Core Structural Connectivity Network 7 and β(w∗, E∗) = (cid:80) e∈E\E∗ w∗(e) E∗ Finally, we proved that there is an admissible solution for CSCN problem. (cid:3) 1 2 (E(cid:48) − 2k) s.t + 2k = = B. Lemma 4. If there is an admissible solution for CSCN problem, then there is an admissible solution for ST problem. 1 2 (E(cid:48)−2k) s.t+2k = B. Proof. Suppose there is an admissible solution for CSCN problem. We prove that there is an admissible solution for ST problem. Let E∗ ⊆ E be such that the graph induced by E∗ is connected, and such that α(w∗, E∗) ≥ s.t+k s.t+2k = A and β(w∗, E∗) ≤= We first prove that {{vi,j, ui} 1 ≤ i ≤ s, 1 ≤ j ≤ t} ⊆ E∗. By Corollary 1, e∈E∗ w∗(e) ≥ s.t + k. By construction of G, the set of edges of weight 1 is {{vi,j, ui} 1 ≤ i ≤ 2} = {{vi,j, ui} 1 ≤ i ≤ s, 1 ≤ j ≤ t}. s, 1 ≤ j ≤ t}, that is {e ∈ E w∗(e) > 1 We get that {{vi,j, ui} 1 ≤ i ≤ s, 1 ≤ j ≤ t} ⊆ E∗ because, otherwise, we we know that E∗ = s.t + 2k. Thus, it necessarily means that(cid:80) would have(cid:80) we would have(cid:80) e∈E∗ w∗(e) < s.t + k. e∈E∗ w∗(e) < s.t + k. Furthermore, every e ∈ E∗ ∩ E(cid:48) is such that w∗(e) = 1 2 . Indeed, otherwise, Finally, since E∗ is an admissible solution for CSCN problem, then it means that the graph induced by the set of edges E∗ ∩ E(cid:48) is connected and is such that for every ui, 1 ≤ i ≤ s, then there is an edge e ∈ E∗ ∩ E(cid:48) that is adjacent to ui. By the previous remark, every edge in E∗ ∩ E(cid:48) has weight 1 2 . Thus, it means that there is E∗ ∩ E(cid:48) is an admissible solution for ST problem considering the (cid:3) graph G(cid:48). Indeed E∗ ∩ E(cid:48) = 2k and so(cid:80) e∈E∗∩E(cid:48) w(e) = k. We are now able to prove the NP-completeness of CSCN problem. Theorem 1. CSCN problem is NP-complete. Proof. The reduction is clearly polynomial. Furthermore, Lemma 3 and Lemma 4 prove the equivalence between CSCN problem and ST problem. Since the deci- sion version of ST problem is NP-complete even if all weights are equal [6], then we obtain the NP-completeness of the decision version of CSCN problem. (cid:3) In Theorem 1 we have proved that CSCN problem is NP-complete. Hence, to be able to solve it in reasonable time we need a relaxation to make it tractable or an approximate algorithm for the complete version. In this article we will propose both. 2.2 Relaxation of the CSCN problem We proved in the previous section that the connectivity constraint is the main reason of the difficulty of the problem. Without it, it becomes tractable. So we solve, in this section, a relaxed version of problem without the connectivity constraint. Then, we use this solution to approximate the full problem. 8 N. Lascano et al. Theorem 2. The decision version of CSCN problem without the connectivity constraint is in P. Algorithm 1 Maximum edges Compute w∗(e) for each e ∈ E Sort(E) for each e ∈ E do E∗ ← E∗ ∪ e if α(w∗, E∗) > A and β(w∗, E∗) < B then return T rue end if end for return F alse (cid:46) sorts edges by w∗ non-increasingly Proof. Algorithm 1, in each step i, defines E∗ as the i maximum weighted edges and tries to use that to fulfill the constraints. Assume that there exists an E∗ that fulfills the constraints. There are two cases: 1) E∗ has the E∗ maximum weighted edges, 2) there are ej ∈ E∗, ek ∈ E \ E∗ such that w∗(ek) ≥ w∗(ej). In 1), Algorithm 1 will find E∗. In 2), let E(cid:48) = (E∗ ∪ {ek}) \ {ej} another subset of E. Then α(w∗, E(cid:48)) = e∈E(cid:48) w∗(e) E(cid:48) = e∈E(cid:48) w∗(e) E∗ ≥ e∈E∗ w∗(e) E∗ = α(w∗, E∗) ≥ A because the edges in E∗ are the same as the ones in E(cid:48) except from one that has a larger weight. For the same reason, (cid:80) (cid:80) (cid:80) (cid:80) ≤ (cid:80) e∈E\E(cid:48) w∗(e) (cid:80) e∈E\E(cid:48) w∗(e) E(cid:48) = E∗ β(w∗, E(cid:48)) = e∈E\E∗ w∗(e) E∗ = β(w∗, E∗) ≤ B. Thus, we found a new subset of E that stills fulfills the constraints. We can do the same process with E(cid:48) (replace an edge with another one of larger weight) iteratively, always getting subsets that fulfills the constraints, until we cannot do this anymore. At that point we will have a subset that has only the maximum E∗ edges and fulfills the constraints. Thus, algorithm 1 will find this subset. We now need to prove algorithm 1 runs in polynomial time in the size of E. The first operation, computing w∗(e) for each e, implies computing the mean and standard deviation for each edge across the population, which can be done in O(N ) per edge (where N is the size of the population). This is O(N ∗E) for all the edges. The second step, sorting, can be done in O(E log E). The main loop runs at most E times, and in each loop it adds an edge to E∗, computes α and β and performs two comparisons. The comparisons can be done in constant time, as the addition to E∗ if we use a linked list of edges to represent it. To compute α and β it is needed to iterate once again E (the part Extracting the Groupwise Core Structural Connectivity Network 9 in E∗ for α, the part in E \ E∗ for β) adding the weights together and then performing two divisions. This can be done in linear time in the size of E, and even quicker (constant time) if we optimize it by keeping the values of α and β across loops and updating them with the weight of the edge that changed sets. Then, algorithm 1 solves the CSCN problem in O(max(E2,E ∗ N ) or in O(E ∗ N ) if a little optimization is used. (cid:3) 2.3 Heuristic approach In Section 2.2 we developed Algorithm 1 to solve the problem of finding the Core Structural Connectivity Network in polynomial time. However, this algorithm does not guarantee a connected result. We solve the original problem, presented in Section 2, by first applying Algorithm 1 and then modifying the resulting core graph G∗ to guarantee its connectedness. This results in an approximate solution for the full problem computable in polynomial time. To extend G∗ into a connected graph we add the necessary edges while de- creasing the minimum possible the objective function f defined in Eq. 6. For this, we use the same approach that Wassermann et al. [10]. Namely, we make a multigraph Gcc with the connected components of G∗ as nodes, complete it with all the possible edges between those connected components, and run a Maximum Spanning Tree algorithm. This selects the edges needed to produce a connected subgraph with the maximum possible weight. For the full details, see Wasser- mann et al. [10]. This way we get a connected subgraph close to the best possible subgraph, which we obtained using Algorithm 1. 3 Experiments and Results Fig. 1. Core structural connectivity network computed by our approach. On the left, we show the adjacency matrix for λ = 0.5, where 48.19% of the connections were included in the CSCN. In the central and right panels, we show the resulting CSCN for λ = 0.9 and 0.99 respectively. The percentage of included connections in the CSCN is 5.99% and 1.27% respectively. We formalized the CSCN problem in section 2 and designed an algorithm to solve it in section 2.3. Now we will asses the performance of our method. For LRLR 10 N. Lascano et al. this, we compare it with the most used [7] and with the recent one [10] in the task of connectivity prediction performance. We use a subset of the HCP500 dataset [9]: all subjects aged 21-40 with complete dMRI protocol, totaling 309. We compute the weighted connectivity matrices between the cortical regions defined by the Desikan atlas [4] as done by Sotiropoulos et al. [9]. Examples of CSCN exctracted with our algorithm at different λ levels are shown in Fig. 1, which was generated using Nilearn [1]. 3.1 Consistency of the Extracted Graph To compare the stability across different algorithms for CSCN, we use an analysis based on Wassermann et al. [10]: we randomly take 500 subsets of 100 subjects each and computed the core graphs for all subsets. We then compute the number of unstable connections: connections that present in at least one core graph but not in all of them. Finally, in Table 1 we report each algorithm's stability: stability of the algorithm (cid:44) 1 − #{unstable connections} #{total connections} . This measure quantifies the CSCN consistency across subsamples. Due to the homogeneity of our sample, we expect the CSCNs obtained by an algorithm to be similar across subsamples. Hence, a stabler algorithm is preferable. 3.2 Predicting Handedness-Specific Connectivity We evaluate performance of the methods by using the generated core graphs as a feature selection for handedness specific connectivity. We use a nested Leave- 3 -Out procedure: the outer loop performs model selection on 1 1 3 of the subjects using the core graph algorithm and the inner loop performs model fitting and prediction using the selected features. Table 1. Stability of the algorithms and amount of features selected by linear regression in the core graph relating the weights with handedness. Our procedure gets more features selected than Gong et al. [7] and Wassermann et al. [10], showing better statistical power. It's also more stable than Gong et al., showing improved consistency. Algorithm Gong et al. 2009 Wassermann et al. 2016 Our approach (λ = 0.50) Features (mean) Features (std) Stability 0.066 0.415 1.042 0.256 0.723 1.269 0.364 0.644 0.528 Specifically, we first take 1 3 subjects randomly and compute the core graph for those subjects using the three different algorithms. Then we add the weights for the selected edges for each subject, and select the features F that are more deter- minant of handedness using a linear least-squares regression and the Bonferroni Extracting the Groupwise Core Structural Connectivity Network 11 Fig. 2. Performance of core network as feature selection for a linear model for handed- ness specific connectivity. We evaluate model prediction (left) and fit (right) for Gong et al. [7] in green, Wassermann et al. [10] in blue and ours, in red. We show the his- tograms from our nested Leave- 1 3 -Out experiment. In both measures, our approach has more frequent lower values than Gong et al., showing a better performance. correction for multiple hypotheses. This experiment is repeated 500 times. We quantify the amount of features that are selected after each procedure, which indicates how useful is the core graph algorithm for selecting the edges related to handedness. We show the results in Table 1. To evaluate the prediction, we randomly take 1 2 of the remaining subjects and fit a linear model on the features F to predict connectivity weights using the handedness of each subject. Finally, we predict the values of the features F from the handedness in the subjects left out. We quantify the quality of the linear model fitting Akaike Information Criterion (AIC) and of the prediction performance with the mean squared error (MSE) of the prediction. For both measures a lower value indicates better performance. The outer loop is performed 500 times and the inner loop 100 times per outer loop, which totals 50,000 experiments. We show the results of this experiments in Fig. 2. 4 Discussion and Conclusion We presented for the first time a polynomial algorithm to extract the core struc- tural connectivity network of a population combining a graph-theoretical ap- proach with statistic relevance of the connections, observing the recent evidence of the structural network topology. Our results show that our algorithm outperforms, in the prediction experi- ment, the most used technique [7] as well as latest approaches [10]. In Table 1 we can see that our algorithm preserves, in average, more connections correlated with the handedness of the subjects. We can also see that despite being less 012345Value1e40.00.20.40.60.81.0Frecuency1e4MSE25002000150010005000Value0123456Frecuency1e3AICWassermann et al. 2016Gong et al. 2009Our approach (λ = 0.50) 12 N. Lascano et al. stable than Wassermann et al.'s it is stabler than Gong et al.'s. Finally, Fig. 2 shows that, in the handedness prediction experiment, our method outperforms Gong et al.'s and Wassermann et al's: the number of cases with lower AIC and MSE is larger in our case. Hence, our CSCN is better as linear model relating connectivity with handedness in terms of model fitting and prediction. In terms of theoretical contributions, we formalized the problem, proved its difficulty and gave a novel algorithm for dealing with it. We then validated our approach by showing its power as feature selector for getting connections related to handedness with 300 real subjects' data. The experiment shows our method performs better than the currently available. Moreover, our method avoids the double dipping problem by not choosing the feature set with hypothesis testing. Acknowledgements Authors acknowledge funding from ERC Advanced Grant agreement No 694665 : CoBCoM - Computational Brain Connectivity Mapping References 1. Abraham, A., Pedregosa, F., Eickenberg, M., Gervais, P., Mueller, A., Kossaifi, J., Gramfort, A., Thirion, B., Varoquaux, G.: Machine learning for neuroimaging with scikit-learn. Frontiers in neuroinformatics 8(February), 14 (2014) 2. Bassett, D.S., Porter, M.A., Wymbs, N.F., Grafton, S.T., Carlson, J.M., Mucha, P.J.: Robust detection of dynamic community structure in networks. Chaos 23(1) (2013) 3. Bullmore, E.T., Sporns, O., Solla, S.A.: Complex brain networks: graph theoretical analysis of structural and functional systems. Nature reviews. Neuroscience 10(3), 186–98 (2009), http://www.ncbi.nlm.nih.gov/pubmed/19190637 4. Desikan, R.S., S´egonne, F., Fischl, B., Quinn, B.T., Dickerson, B.C., Blacker, D., Buckner, R.L., Dale, A.M., Maguire, R.P., Hyman, B.T., Albert, M.S., Killiany, R.J.: An automated labeling system for subdividing the human cerebral cortex on MRI scans into gyral based regions of interest. NeuroImage 31(3), 968–980 (2006) 5. Donahue, C.J., Sotiropoulos, S.N., Jbabdi, S., Hernandez-Fernandez, M., Behrens, T.E., Dyrby, T.B., Coalson, T., Kennedy, H., Knoblauch, K., Van Essen, D.C., Glasser, M.F.: Using Diffusion Tractography to Predict Cortical Connection Strength and Distance: A Quantitative Comparison with Tracers in the Monkey. Journal of Neuroscience 36(25), 6758–6770 (jun 2016), http://www.jneurosci. org/cgi/doi/10.1523/JNEUROSCI.0493-16.2016 6. Garey, M.R., Johnson, D.S.: Computers and Intractability: A Guide to the Theory of NP-Completeness. W. H. Freeman & Co., New York, NY, USA (1979) 7. Gong, G., He, Y., Concha, L., Lebel, C., Gross, D.W., Evans, A.C., Beaulieu, C.: Mapping anatomical connectivity patterns of human cerebral cortex using in vivo diffusion tensor imaging tractography. Cerebral Cortex 19(3), 524–536 (2009) 8. Kriegeskorte, N., Simmons, W.K., Bellgowan, P.S.F., Baker, C.I.: Circular analysis in systems neuroscience: the dangers of double dipping. Nature Neuroscience (2009) 9. Sotiropoulos, S.N., Jbabdi, S., Xu, J., Andersson, J.L., Moeller, S., Auerbach, E.J., Glasser, M.F., Hernandez, M., Sapiro, G., Jenkinson, M., Feinberg, D.A., Yacoub, E., Lenglet, C., Van Essen, D.C., Ugurbil, K., Behrens, T.E.: Advances in diffusion MRI acquisition and processing in the Human Connectome Project. NeuroImage 80(3), 125–143 (oct 2013) 10. Wassermann, D., Mazauric, D., Gallardo Diez, G.A., Deriche, R.: Extracting the Core Structural Connectivity Network: Guaranteeing Network Connectedness Through a Graph-Theoretical Approach. In: MICCAI 2016 (2016)
1908.03514
1
1908
2019-08-09T16:13:12
A practical guide to methodological considerations in the controllability of structural brain networks
[ "q-bio.NC", "q-bio.QM" ]
Predicting how the brain can be driven to specific states by means of internal or external control requires a fundamental understanding of the relationship between neural connectivity and activity. Network control theory is a powerful tool from the physical and engineering sciences that can provide insights regarding that relationship; it formalizes the study of how the dynamics of a complex system can arise from its underlying structure of interconnected units. Given the recent use of network control theory in neuroscience, it is now timely to offer a practical guide to methodological considerations in the controllability of structural brain networks. Here we provide a systematic overview of the framework, examine the impact of modeling choices on frequently studied control metrics, and suggest potentially useful theoretical extensions. We ground our discussions, numerical demonstrations, and theoretical advances in a dataset of high-resolution diffusion imaging with 730 diffusion directions acquired over approximately 1 hour of scanning from ten healthy young adults. Following a didactic introduction of the theory, we probe how a selection of modeling choices affects four common statistics: average controllability, modal controllability, minimum control energy, and optimal control energy. Next, we extend the current state of the art in two ways: first, by developing an alternative measure of structural connectivity that accounts for radial propagation of activity through abutting tissue, and second, by defining a complementary metric quantifying the complexity of the energy landscape of a system. We close with specific modeling recommendations and a discussion of methodological constraints.
q-bio.NC
q-bio
A practical guide to methodological considerations in the controllability of structural brain networks Teresa M. Karrera,b, Jason Z. Kimb,*, Jennifer Stisoc,*, Ari E. Kahnc, Fabio Pasqualettid, Ute Habela,e,f, and Danielle S. Bassettb,g,h,i,j,k,l aDepartment of Psychiatry, Psychotherapy and Psychosomatics, Faculty of Medicine, RWTH Aachen, Germany. bDepartment of Bioengineering, School of Engineering & Applied Science, University of Pennsylvania, Philadelphia, PA 19104, USA. cDepartment of Neuroscience, Perelman School of Medicine, University of Pennsylvania, Philadelphia, PA 19104, USA. dDepartment of Mechanical Engineering, University of California, Riverside, CA 92521, USA. eJARA - Translational Brain Medicine, Aachen, Germany. fInstitute of Neuroscience and Medicine: JARA-Institute Brain Structure Function Relationship (INM 10), Research Center Jlich, gDepartment of Physics and Astronomy, College of Arts & Sciences, University of Pennsylvania, Philadelphia, PA 19104, USA. hDepartment of Neurology, Perelman School of Medicine, University of Pennsylvania, Philadelphia, PA 19104, USA. iDepartment of Psychiatry, Perelman School of Medicine, University of Pennsylvania, Philadelphia, PA 19104, USA. jDepartment of Electrical and Systems Engineering, School of Engineering & Applied Science, University of Pennsylvania, Jlich, Germany. Philadelphia, PA 19104, USA. kSanta Fe Institute, Santa Fe, NM 87501, USA. lTo whom correspondence should be addressed: [email protected] *These two authors contributed equally. Abstract Predicting how the brain can be driven to specific states by means of internal or external control requires a fundamental understanding of the relationship between neural connectivity and activity. Network control theory is a powerful tool from the physical and engineering sciences that can provide insights regarding that relationship; it formalizes the study of how the dynamics of a complex system can arise from its underlying structure of interconnected units. Given the recent use of network control theory in neuroscience, it is now timely to offer a practical guide to methodological considerations in the controllability of structural brain networks. Here we provide a systematic overview of the framework, examine the impact of modeling choices on frequently studied control metrics, and suggest potentially useful theoretical extensions. We ground our discussions, numerical demonstrations, and theoretical advances in a dataset of high-resolution diffusion imaging with 730 diffusion directions acquired over approximately 1 hour of scanning from ten healthy young adults. Following a didactic introduction of the theory, we probe how a selection of modeling choices affects four common statistics: average controllability, modal controllability, minimum control energy, and optimal control energy. Next, we extend the current state of the art in two ways: first, by developing an alternative measure of structural connectivity that accounts for radial propagation of activity through abutting tissue, and second, by defining a complementary metric quantifying the complexity of the energy landscape of a system. We close with specific modeling recommendations and a discussion of methodological constraints. Our hope is that this accessible account will inspire the neuroimaging community to more fully exploit the potential of network control theory in tackling pressing questions in cognitive, developmental, and clinical neuroscience. Keywords: network neuroscience, control theory, structural connectivity, diffusion imaging. 1 1 Introduction The brain is a complex system of interconnected units that dynamically transitions through diverse ac- tivation states supporting cognitive function [1]. Understanding the mechanisms and processes that give rise to these trajectories through state space is crucial for intervening in disease to restore cognitive functioning [2]. One relevant factor enabling such rich neural dynamics is the network architecture of the underlying structural substrate [3 -- 5]. Yet, the exact mechanisms by which the physical architecture of the brain both supports and constrains its function remain largely unknown [6 -- 8]. Recent advances in network control theory offer a formal means to study how the temporal dynam- ics of a complex system emerges from its underlying network structure [9 -- 11]. Applying this theory to the brain requires that one first builds a network model in which brain regions (nodes) are anatomically connected to one another (edges) [12, 13]. The state of the brain network system is then reflected in the pattern of neurophysiological activity across network nodes, and state trajectories represent the temporal sequence of brain states that the system traverses [14, 15]. With definitions of the network and its state in hand, we can consider the problem of network controllability, which in essence amounts to asking how the system can be driven to specific target states by means of internal or external control input [16]. In the context of the brain, such input can intuitively take the form of electrical stimulation [17 -- 21], task modulation [22 -- 24], or other perturbations from the world or from different portions of the body [25, 26]. Practically, network control theory and its associated toolkit enables us to study the general role of brain regions in controlling neural dynamics in diverse scales and species [27 -- 30], and in both health [31, 32] and disease [33, 34] or injury [35]. Moreover, the approach can be used to determine the patterns of input required to induce specific state transitions necessary for behavior [18, 22, 35, 36]. Network control theory offers three primary advantages over traditional approaches to the study of brain network function. First, the multi-modal nature of the theoretical framework explicitly enforces a simultaneous study of brain structure and function, in contrast to approaches that characterize each separately and then assess statistical covariance. Second, network control theory exceeds the often purely descriptive approach of network science [37 -- 39] by building a generative model parameterized by both a network's spatial features and its temporal features [40]. The model then offers predictions of the brain's response to both endogenous and exogenous input signals. In the case of the former, the model could hypothetically prove useful in understanding how the brain enacts cognitive control to reach task-relevant cognitive states [22, 24, 31]. In the context of the latter, the model could similarly prove useful in in- forming neuromodulation for the treatment of neurological and psychiatric disorders [40]. Third, initial studies applying network control theory in neuroscience demonstrate that network controllability is a useful marker of brain dynamics, from quantifying the capacity of different brain regions to alter whole- brain dynamics [31], over demonstrating that this capacity grows with development [41], to finding that controllability is linked to executive functioning [22, 24, 42]. Moreover, applications of the theory to data collected during invasive neuromodulation regimens demonstrate the utility of the theory in predicting response to electrical stimulation in practice [18, 21] and in theory [17]. In light of the promising applicability of network control theory in neuroscience, we wish to provide a systematic overview of how the framework can be used to study the controllability of neural dynamics. This primer is constructed so as to offer neuroscientists some basic intuitions regarding the foundational concepts, and to guide them through the necessary prerequisites and considerations. For a more technical 2 introduction that nevertheless remains heavily motivated by neuroscience, we refer the interested reader to [43, 44]; and for further information about the underlying mathematics (which remains agnostic to the application domain), we refer the reader to [9, 11]. Because the application of network control theory can be formulated in several ways, we systematically probe how diverging theoretical assumptions and possible modeling choices influence controllability metrics and the estimated energy of state transitions. For example, we consider discrete and continuous time systems, methods for system stabilization, the time horizon for control, and the set of control nodes. We complement these studies with specific rec- ommendations for best practices, which depend in no small part upon the nature of the neuroscientific question being investigated. To further stimulate research in this exciting field, we suggest a few useful extensions of the theoretical framework, such as alternative estimates of structural connectivity and a complementary metric that quantifies the complexity of the energy landscape. 2 Theoretical framework 2.1 Network control theory The core of the theoretical framework is the structural network of neurons (or larger neural units) in the brain that allows the activity of a brain region to diffuse and change the activity of connected brain regions (Fig. 1A). Here, we introduce a mathematical model that describes the natural dynamics of a complex linear system (Fig. 1B). Formally, the temporal evolution of network activity is modeled as a linear function of its connectivity: x = Ax(t), (1) where x(t) is a vector of size N × 1 that represents the state of the system. Here we operationalize the system's state to reflect the magnitude of the neurophysiological activity of the N brain regions at a single point in time. Over time, x(t) denotes the state trajectory, which is the temporal sequence of states or activity patterns that is traversed by the system. The adjacency matrix A is of size N × N , and denotes the relationships between the system elements. Here, we operationalize that relation as the structural connectivity between each pair of brain regions. Next, we extend this model to account for controlled dynamics, which occur when the brain is induced to deviate from its natural trajectory by the injection of internal or external input signals (Fig. 1C). In this case, the temporal dynamics of a system additionally depends on the control energy injected into a set of nodes across time x = Ax(t) + Bκuκ(t). (2) Here, Bκ is a matrix of size N × m that denotes the set of m control nodes or brain regions into which we wish to inject inputs. This matrix consists of m indicator vectors, each having set only the i-th element to 1, corresponding to a control node. If we control all brain regions, Bκ corresponds to the N × N identity matrix with ones on the diagonal and zeros elsewhere. If we control only a single brain region i, Bκ reduces to a single N × 1 vector with a one in the i-th element and zeros elsewhere. The term uκ(t) is a vector of m functions of size m × 1 denoting the control input, which is the amount of input injected into each of the m control nodes at each time point t. Over time, uκ(t) denotes the injected control input over time. For the interested reader, we wish to provide a few mathematical intuitions that might facilitate a deeper understanding of the presented concepts. By Eq. 1, the structure of the network determines its dynamic evolution over time. Mathematically, the structural connectivity matrix A serves as linear operator that 3 maps each state, x, to the rate of change from that state, x. This linear transformation can be described in terms of the evolutionary modes of the system consisting of the N eigenvectors of A and their associ- ated eigenvalues (Fig. 1D). Each eigenvector of A can be imagined as an axis of the linear transformation which remains invariant over time. Thus, the eigenvectors reflect directions in the state space along which the system independently moves, each characterized by a specific pattern of brain region activity. Each eigenvalue, in turn, determines the rate of growth or decay along its associated eigenvector; that is, each eigenvalue determines how slow or fast the system grows or decays in the direction defined by the eigen- vector. Thus, the eigenvalues control the temporal persistence of the set of supported modes of activity. Especially for the interpretation of results, it is important to keep in mind that the dynamic model is relatively simple and relies on the assumptions of linearity, time invariance, and freedom from noise. Linearity implies that the system evolves linearly over time which is not an accurate reflection of extended dynamics in most neural processes. However, it has been shown that non-linear dynamics can be locally approximated by linear dynamics [45, 46]. Time invariance implies that the systems response does not depend on the time point because both the structural network A and the control set Bκ are constant over time. This assumption likely holds true for short time scales but could be challenged by long-term structural reorganization, which has been observed across development and adulthood [41, 47, 48]. Free- dom from noise implies that all properties of signal propagation are accounted for deterministically by the model. Yet, noise is a feature of neural signals at both small [49 -- 51] and large time scales [52, 53]. Nevertheless, it is customary and reasonable when first developing a mathematical model of a complex system to consider the salient features of the model that do not depend on noise [44, 54, 55]. 2.2 Prerequisites The core of the dynamic model is the network structure that enables activity changes within a particular brain region to diffuse and induce state changes in connected regions. Thus, the first step is to build a structural connectivity network by defining the weighted adjacency matrix A (Fig. 1A). The structural network of human and nonhuman animals can be modeled using a range of spatial scales of neural units and physical links between them [13]. Here, we focus on the construction of the human connectome which requires (i) a brain parcellation that defines the N nodes of the network and (ii) diffusion imaging data that define the strength of structural connectivity Ai,j between two brain regions i and j. To avoid self-loops in the system, we set the diagonal of A to zero. As a further practical note, the sparse nature of human connectomes typically does not require any thresholding of the matrix. The next step is to pick a time-system that best reflects the neural dynamics under study. Here, we consider two options: discrete and continuous. A discrete-time system assumes that the system evolves in discrete time steps whereas a continuous-time system models continuously changing dynamics. Neural processes can often not be clearly assigned to one of these categories. The spatial scale of the neural unit under study is one factor that could guide this choice [44]. Smaller spatial scales such as single neurons could be modeled as either firing or not, and the accompanying uniform delays in neural activity could be better approximated by discrete-time systems. Larger neural units such as brain regions usually contain more neurons. The ensuing more heterogeneous timing of neural state changes might be bet- ter represented by continuous-time systems. Depending on the choice, the modeled dynamics can differ substantially because of their distinct mathematical implementation. More concretely, discrete-time dy- namics rely on difference equations whereas continuous-time systems are based on differential equations. Note that we exclusively present formulas for continuous-time systems in the main text; the discrete-time 4 Figure 1: Schematics of network control theory and relevant concepts. (A) Structural brain network construction. Brain atlas and diffusion imaging data define the nodes and edges of the structural connectivity matrix A. (B) Natural dynamics of the brain. The temporal evolution of brain states, such as the magnitude of neurophysiological activity across brain regions, is modeled as a linear function of brain structure. The area under the curve illustrates the impulse response and thus, average controllability of brain region #4. (C) Controlled dynamics. The state trajectory additionally depends on control input injected into the system. The control input matrix Bκ determines the nodes into which a control signal uκ is injected (yellow flash) over time. The area under the curve of the control energy signals corresponds to the control energy required by the given state transition. (D) Activity modes of a system. The structural connectivity matrix A can be decomposed into N eigenvectors and eigenvalues that determine the system's dynamics. Eigenvectors determine the supported modes of activity; eigenvalues determine the rate of decline of their associated mode. Brain region i's controllability vi,j of mode j corresponds to a projection of the jth eigenvector onto the dimension spanned by brain region i. (E) Control energies and controllability metrics. (Left) Control energy for specific state transitions. Here we illustrate the minimum control energy required to drive the brain from a specific initial state to a specific target state using a particular control node set. The optimal control energy additionally constrains the size of the state trajectory. (Right) Control strategies potentially examining all possible state transitions (dashed arrows). Average controllability has been previously described as a brain region's ability to control nearby states that require little energy. Modal controllability has been previously described as a brain region's ability to control distant states that require more energy. (F) Controllability metrics and control energies can be relevant on an individual and regional level. To examine both levels separately, we will summarize statistics across brain regions and individuals, respectively. For consistency between two parameter choices such as discrete- and continuous-time systems, we will calculate the Pearson correlation between individual (regional) values extracted from one parameter choice and those extracted from the second parameter choice. G) Complexity of the energy landscape. The landscape of possible minimum control energy trajectories is determined by the eigenvalues of the inverse of the controllability Gramian Wκ,T . We used the variability of the eigenvalues to quantify the heterogeneity of the energy landscape. Abbreviations: IQR, interquartile range. versions can be found in the Supplementary Formulas. 5 The third step is to choose a method to stabilize the system to avoid its infinite growth over time. Because extremely large brain states are neurobiologically implausible, we normalize the system such that it either approaches the largest supported mode of activity or goes to zero over time: Anorm = A λ(A)max + c − I (3) Here, I denotes the identity matrix of size N × N , and λ(A)max denotes the largest eigenvalue of the system. To normalize the system, we must specify the parameter c, which determines the rate of stabi- lization of the system. If c = 0, the largest mode of activity is stable and all other modes decay; thus, the system approaches the largest mode over time. If c > 0, such as the commonly used choice c = 1, all modes decay; thus, the system goes to zero over time. As will become clear in the next section, the latter variant can be especially useful for the computation of average controllability in infinite time as well as modal controllability due to its mathematical definition. 2.3 Optimal control energy To quantify the degree of controllability of a network, we consider an optimal control problem to steer the network from a specific initial state x(0) = x0 to a specific target state x(T ) = xT over the time horizon T while minimizing a combination of both the length of the state trajectory and the required control energy [35, 56, 57]. Formally, we consider the problem u(t)∗ κ = argmin uκ J(uκ) = argmin uκ 0 ((xT − x(t))(cid:62)(xT − x(t)) + ρuκ(t)(cid:62)uκ(t))dt, (4) where the parameter ρ determines the relative weighting between the costs associated with the length of the state trajectory and input energy. We use the cost function J(u(t)∗ κ) to find the unique optimal control input u(t)∗ κ which allows us to calculate the optimal control energy (Fig. 1E) required by a single brain region i (Fig. 1C): (cid:90) T (cid:90) T i (t)(cid:107)2 (cid:90) T (cid:107)u∗ 0 2dt, E∗ i = N(cid:88) i=1 (5) (6) and in total E∗ = E∗ i = u∗ κ(t)(cid:62)u∗ κ(t)dt. 0 To calculate optimal control energy, we must specify an initial brain state x0 and a target brain state xT by assigning each brain region an initial and target activity level. If available, we can extract regional activity values directly from functional neuroimaging data such as electrocorticography or magnetic res- onance imaging [18, 58, 59], or we can use model-based estimates of task-related activation such as β values from a general linear model [42]. Otherwise, we can also model brain states by artificially defining a subset of nodes to be active, such as brain regions belonging to the same cognitive system [22, 35, 36]. Additionally, we must specify the control set Bκ, a set of brain regions into which we wish to inject signals. Theoretically, this choice can vary from controlling a single region to controlling the full brain. The choice of small- to medium-sized control sets, however, can lead to large numerical instabilities that accumulate and bias the results. As a rule of thumb, it is advisable to ensure that the numerical error does not exceed 10−6. To reduce the numerical error of the calculation, we can also define a relaxed control set Bκ by allowing large control input to control regions and small, random inputs to all other brain regions [18]. We must also specify the time horizon T over which the control input is effective. For 6 pragmatic reasons such as the potential translation to real external brain stimulation, the time horizon is usually set to finite time. Note that the time horizon is measured in arbitrary units even if brain states are defined by functional imaging data. Finally, we must specify the time step dt, which should be small enough to sufficiently approximate the dynamics [35]; a reasonable choice is dt = 0.001. The cost function J is motivated by the fact that biological systems might constrain the features of the traversed states, such as their type, diversity, or magnitude. Transitioning through states not too far away from the target state is supposed to avoid extremely large and thus neurobiologically implausible brain state transitions. In the case where no specific assumptions are made on the relative importance of the two constraints and where both the distance and energy values are of a comparable scale, an equal weighting of ρ = 1 is a reasonable choice. Depending on our neurobiological assumptions, we can also define alternative cost functions and potentially restrict them to a subset of brain regions [22]. 2.4 Minimum control energy A specific and commonly used subform of optimal control energy is obtained by letting ρ → ∞ in (4), so that the cost function J accounts only for the energy of the control input to steer the network from an initial state x(0) = x0 to a target state x(T ) = xT . Thus, we call this metric minimum control energy (Fig. 1E). To compute the minimum control energy for a given network, it is convenient to define the controllability Gramian as (cid:90) T 0 Wκ,T = eAtBκB(cid:62) κ eA(cid:62)tdt. (7) The eigenvalues of Wκ,T can be used to answer several questions regarding the controllability of a net- work. First, if the smallest eigenvalue of Wκ,T is zero, then the network is not controllable. That is, there exist final states xT that cannot be reached by any control input, independent of its energy. Second, the magnitude of the smallest eigenvalue of Wκ,T is inversely proportional to the largest energy needed to reach a final state. That is, there exists a final state xT that can be reached only using inputs whose energy is at least proportional to the inverse of the smallest eigenvalue of Wκ,T . The foundational papers [31, 60] have shown that brain networks are controllable from any single region; that is, the smallest eigenvalue of Wκ,T is greater than zero. However, brain networks require very large control energy; that is, the smallest eigenvalue of Wκ,T can be extremely small. It should also be noted that the computation of the smallest eigenvalue of Wκ,T tends to be numerically difficult, which motivates the next metric. 2.5 Average controllability Apart from examining specific state transitions, the theoretical framework also allows us to ask questions regarding the general role of brain regions in controlling neural dynamics. A third metric is obtained by measuring the average input energy required to drive the system via a specified set of control nodes to all possible target states xT with unit norm [61, 62]. Following the above discussion, this metric equals T race(W −1 κ,T ) where the inverse of the controllability Gramian is a map from target states to control energy. To avoid numerical difficulties when controlling only a few nodes in very large systems, the met- ric is often approximated as T race(Wκ,T ). Using this last form, our definition of average controllability measures the ability of a network to amplify and spread control inputs, rather than being associated with the problem of steering the network state from x0 to xT . More concretely, average controllability quantifies the energy of the impulse response of a system, which describes how a system naturally evolves over time from some initial condition [9]. Starting from an exclusive activation of the specified control regions, we observe the brain's natural response (Fig. 1B). The larger and more variable this natural 7 response, the more states can be reached with low energy input by controlling this specific set of brain regions. In prior work, average controllability was also intuitively described as the ability of a set of control nodes to drive the system to easily reachable, nearby states (Fig. 1E) [31]. To calculate average controllability, we must specify the time horizon T , which is the time period over which we wish to observe the impulse response of the system. Note that the units of the time horizon depend on the units of A. To observe the complete impulse response, we often assume infinite time. Fur- thermore, we must determine the control set Bκ, which is the set of brain regions into which control input can be injected. Even if the control set can comprise multiple, and even all nodes, average controllability is often examined for individual brain regions to enable comparison to another single-node metric: most commonly, modal controllability. 2.6 Modal controllability Lastly, we introduce the metric modal controllability, which was previously described as the ability of a single node to drive the system to distant, more difficult-to-reach states (Fig. 1E) [31]. The controllability metric is obtained directly from the eigenvalues and eigenvectors of the network weighted adjacency matrix. In particular, we use N(cid:88) φi = (1 − (eλj (A)))v2 ij, (8) j=1 as a scaled summary of node i's ability to control all N modes of the network [11]. To calculate modal controllability, we are not required to specify any parameters except the symmetric adjacency matrix A. This metric capitalizes on information housed in the modes of A, as summarized in the eigenvalues λj and the matrix of normalized eigenvectors V = [vi,j]. Entry vi,j is a measure of the controllability of mode λj(A) from node i that geometrically corresponds to projecting node i onto the eigenvector j (Fig. 1D) [9, 63]. According to this heuristic, the larger the magnitude of the projection, the higher the ability of node i to control mode j. The metric summarizes this notion across all modes, and then scales them by their rate of decline as determined by the eigenvalues. This weighting emphasizes especially fast decaying modes which might on average be more difficult to control because the injected control energy only has a short-term impact. We note that modal controllability is exclusively formalized for symmetric matrices whereas all of the other definitions that we present can be naturally extended to directed networks. For completeness, we note that boundary controllability is another controllability metric used in the literature and measures the ability of a brain region to integrate information between network commu- nities [11, 31]. However, because the metric is less commonly used, we will not discuss it further in this work. 3 Materials and methods 3.1 Acquisition of diffusion imaging data High resolution anatomical brain images were collected from 10 healthy young adults (23.9 ± 3.6 years; 20-31 years; 70% female). The participants underwent a 53:24 minute diffusion spectrum imaging (DSI) scan with 730 diffusion directions (maximum b-value = 5010s/mm2, 21 b = 0 images, TR = 4300ms, TE = 102ms, matrix size = 144×144, field of view = 260 × 260mm2, slice number = 87, resolution = 1.8×1.8×1.8mm3, multi-band acceleration factor = 3). Additionally, T1-weighted images were obtained 8 using an MPRAGE sequence (TR = 2500ms, TE = 2.18ms, flip angle = 7 degrees, slice number = 208, slice thickness = 0.9mm). Both scans were acquired on a Siemens Magnetom Prisma 3 Tesla scanner with a 64-channel head coil. The study was approved by the Institutional Review Board of the University of Pennsylvania and all participants provided informed consent in writing. 3.2 Preprocessing of diffusion imaging data As previously described in more detail [30], the individual DSI scans were skull-stripped, realigned, and motion-corrected using an improved average b=0 reference image. The preprocessing was implemented in nipype [64] using the Advanced Normalization Tools (ANTs, [65]) for image registration. We quanti- fied the diffusion at different orientations in each voxel using the generalized q-sampling reconstruction method [66] in DSI Studio (dsi-studio.labsolver.org). Based on the derived quantitative anisotropy values, we performed deterministic tractography across the whole-brain [67]. For each participant, we generated 1,000,000 streamlines with a maximum length of 500mm [68] and a maximum turning angle of 35 degrees [69]. 3.3 Construction of structural brain networks Based on the diffusion imaging data, we constructed a structural brain network for each participant. Consistent with previous work [31, 35, 36, 41], we defined nodes of the network as brain regions according to the 234-node Lausanne atlas (excluding brainstem) [70]. For this purpose, the Lausanne parcels were dilated by 4mm so that the parcels reached down into the white matter enough to ensure accurate sam- pling of underlying fibers. In the process of dilation, some voxels were assigned to two or more regions of interest; to eradicate this redundancy, we assigned each voxel to the mode of its neighbors [71]. After warping the parcellation into the subject's diffusion space, we quantified the edges of the network as total streamline count connecting a pair of brain regions, corrected for their volume. Overall, we constructed a 233×233 sparse, weighted, and undirected adjacency matrix for each participant with the number of interregional streamlines representing structural connectivity. 3.4 Mapping to cognitive systems To define neurobiologically meaningful brain states, we capitalized on an established functional brain atlas [72]. By clustering the resting state functional magnetic resonance imaging data of 1000 healthy adults, Yeo et al. identified seven cognitive systems, each consisting of a set of distributed brain regions that are functionally coupled [72]. The functional parcellation comprises visual (VIS), somatomotor (SOM), dorsal attention (DOR), ventral attention (VEN), limbic (LIM), frontoparietal control (FPC), and default mode (DM) systems. To link the functional and anatomical atlases, we mapped each brain region to the cognitive system with the highest spatial overlap as reported previously [22, 48]. More concretely, each Lausanne parcel was assigned to the cognitive system that was most frequently associated with its voxels as defined by the purity index. Subcortical regions were summarized in an eighth, subcortical system (SC). 3.5 Probing different modeling choices We used the structural connectivity matrices of our sample to probe the impact of several modeling choices on average and modal controllability, and on minimum and optimal control energy. In our anal- yses, we systematically varied one parameter at a time while keeping all other parameters constant. Constant modeling choices were guided by the modeling choices most commonly used in the literature [31, 35, 36, 41]. Concretely, we employed a simplified noise-free linear continuous-time and time-invariant network model, stabilized using c = 1. When estimating average controllability, we set the time horizon 9 T to infinite time. When estimating control energies, we used T = 3, approximated by 1000 time steps. When the system matrix A is stable, the controllability Gramian equation converges as T approaches infinity. In this case, the Gramian can be computed algebraically by solving the Lyapunov equation. Motivated by the questions most relevant to each approach, we calculated average and modal controlla- bility for each brain region based on single-node control sets, whereas control energies were based on full brain control. To examine control energies for transitions between previously defined functional systems, we simulated state transitions from an initially active default mode system to the activation of six differ- ent cognitive systems representing the target states [72]. For each specific brain state, regions belonging to the activated cognitive system were set to one, whereas all other brain regions were set to zero. Except for the section on full versus partial control, we averaged across the examined state transitions. For op- timal control energy, we set the relative energy weight ρ = 1. In the restricted set of state transitions we investigated, minimum and optimal control energy yielded highly similar results. To avoid redundancy, we report the results on optimal control energy in the Supplementary Results (SFig. 1-6). Nevertheless, we point out deviating results of optimal control energy in the main text. 3.6 Examining metrics on an individual and regional level Since network control theory can be utilized to examine controllability differences in both individuals and brain regions, we separately studied the metrics on an individual and regional level. For this purpose, we summarized average controllability, modal controllability, and minimum control energy across either brain regions or individuals to subsequently investigate individual and regional values, respectively (Fig. 1F). To estimate the consistency of a metric across parameter choices on an individual (regional) level, we first summarized the metric across brain regions (individuals) and then computed the Pearson correlation between the individual (regional) values obtained with one modeling choice and the individual (regional) values obtained with the second modeling choice; for example, we compare discrete- and continuous-time systems, and we compare two different parameter choices of time horizon T . 3.7 Construction of spatial adjacency network In addition to diffusing along white matter fibers, neural signals could potentially also diffuse between spatially adjacent brain regions. In other words, physical contact between two regions can be seen as a form of structural connectivity. To examine this complementary measure of structural connectivity, we generated brain networks, S, based on the amount of shared neighborhood between two brain regions. We defined the edges of S as the number of face-touching voxels between two parcels of the Lausanne-atlas warped into subject space. In addition to studying each structural matrix separately, we also exploit the combined information of both measures by constructing the matrix AS as an average of A and S. Because both diffusion and adjacency measures are expressed in arbitrary units and the actual scaling might impact controllability metrics, we scaled S and AS to the range of A. We tested the effects of the structural connectivity types and their binarized version in a repeated measures ANOVA with two within-subject factors. To ensure that the effects of matrix type and binarization were not exclusively based on different edge weight distributions [73], we verified the results by sampling edge weights of S and AS from the distribution of A while preserving their rank order. 3.8 Controllability of fast and slow dynamics We capitalized on the concept of modal controllability to probe the ability of a brain region to control a specific set of temporal dynamics such as fast and slow modes [18, 74]. Instead of summarizing across 10 all modes that a system supports, we restricted the calculation of modal controllability to a subset of fastest (slowest) modes. We define transient (persistent) modal controllability as the ability of a brain region to control fast (slow) modes. The temporal dynamics of modes are determined by the magnitude of their eigenvalues. In continuous-time systems, large (small) eigenvalues relate to quickly (slowly) de- caying modes. The lack of a formal definition of fast and slow dynamics requires the choice of a threshold that specifies the subset of modes (Fig. 1D). We systematically probed the influence of threshold on a brain region's ability to control different temporal dynamics by calculating transient and persistent modal controllability using the 10%, 20%, 30%, 40%, and 50% fastest and slowest modes. To disentangle these overlapping control tasks, we additionally summarized the ability of each brain region to control a specific interval of modes using the unscaled eigenvector matrix V . For this purpose, we separated V into both 10 intervals and 2 intervals; this separation enabled a comparison to persistent and transient modal controllability based on a cut-off of 10% and 50%, respectively. 3.9 Definition of complexity of the energy landscape The control trajectories from any initial state to any target state span the energy landscape of a dynamic system. The heterogeneity of the minimum control energy landscape is determined by the eigenvalues of the inverse of the controllability Gramian [9]. Note that we assume independent control from all brain regions because the inverse of the Gramian is often ill-conditioned for small control sets [31]. We capitalized on the variability of the eigenvalues to quantify the complexity of the minimum control energy landscape of a brain network, that is how the magnitude of the minimum control energy varies across all possible state transitions (Fig. 1G). To account for the observed skewness of the distribution, we adopted the interquartile range as a measure of variability. Formally, we define the complexity of the energy landscape as the difference between the 75th and 25th percentile of the eigenvalue distribution of the inverted controllability Gramian Cκ,T = P75(λW −1 κ,T ) − P25(λW −1 κ,T ). (9) We calculated the complexity of the energy landscape for each participant based on an infinite-time controllability Gramian. Then, we tested the complexity of the energy landscape of the brain network against three null models preserving distinct network characteristics. The topological null model preserved degree and strength distribution by iteratively switching connections between randomly selected edge pairs and subsequently associating the connections with the empirically observed edge weights [75]. The spatial null model preserved the relationship between Euclidean distance on the edge weights by adding the initially removed distance effects to the randomly rewired graph [76]. The combined null model preserved both the strength distribution and spatial embedding of the brain networks by approximating the observed strength distributions and effects of Euclidean distance on the edge weights [76]. Overall, we generated 1000 random instantiations of each null model. 4 Results In the application of network control theory, we can rely on different neurobiological assumptions that are reflected in our modeling decisions. We begin with an examination of the impact of different modeling choices, before investigating several proposed model extensions. 11 4.1 Consistency across time systems When examining how the brains architecture gives rise to its complex dynamics by means of network control theory, one of the first modeling steps represents the type of the dynamic model. We can either assume that the neural dynamics evolve in discrete time steps or continuously. In light of potentially distinct dynamics of discrete- and continuous-time systems, we initially examined the consistency of min- imum control energy, average controllability, and modal controllability across time systems. For this purpose, we calculated the Pearson correlation of each metric between discrete- and continuous-time sys- tems, separately summarized across brain regions and individuals, and -- if applicable -- for different time horizons T (Fig. 2A). Average controllability showed a high consistency across time systems (individual level: rmin = 0.80, p = 5 × 10−3; regional level rmin = 0.99, p = 5 × 10−221), particularly for time horizons close to zero or infinity. Likewise, modal controllability demonstrated a high consistency across time systems (individual level: r = 0.99, p = 3 × 10−11; regional level r = 1.0, p = 2 × 10−16). Minimum control energy, however, was less consistent across time systems (individual level: rmin = 0.77, p = 0.01; regional level rmin = 0.33, p = 2 × 10−7), particularly for short time horizons. The observed results are in line with theoretical considerations that suggest a convergence of discrete- and continuous-time systems for infinite time. Overall, the consistency across discrete- and continuous-time systems was high but depended on the metric, the observation level, and the chosen time horizon. Figure 2: Consistency of metrics across time. (A) Consistency of average controllability and minimum control energy across time systems. Pearson correlation coefficient between a given metric estimated for discrete- versus continuous-time systems across a range of time horizons T . (Left) Average controllability; (Right) minimum control energy. (B) Consistency of average controllability and minimum control energy in a continuous-time system for any two choices of time horizon. Heat maps depict correlation matrix of different time horizons. Each heat map entry corresponds to the Pearson correlation of a metric based on two different time horizon choices. (Top) Pearson correlation of individual metrics averaged across brain regions. (Bottom) Pearson correlation of regional metrics averaged across participants. From these results, we deduce that average controllability and minimum control energy differ qualitatively for discrete- versus continuous-time systems when comparing estimates from short time horizons versus from longer time horizons. 12 4.2 Consistency across time horizons Network control theory might lend itself particularly well to evaluate how local perturbations of the brain, for instance elicited by deep brain stimulation or transcranial magnetic stimulation, affect whole brain dynamics. In such a setting we might be interested in assessing different temporal scales of brain stimu- lation such as the effect of stimulation in the short term or in the long term. This question prompts the examination of the time horizon of the injected signal as another early modeling decision. We addressed this question by quantifying the Pearson correlation between values estimated for one time horizon T and for another time horizon T (cid:48), separately averaged across brain regions or across individuals (Fig. 2B). We first noted that the time horizon affected the scaling of the metrics (SFig. 1A). More specifically, aver- age controllability monotonically increased in magnitude with larger time horizons because we observed the impulse response of the system for a longer time interval. Minimum control energy monotonically decreased with larger time horizons; this relation is intuitive when we consider the fact that longer time horizons allow the system to capitalize on its own natural dynamics, thereby demanding less exogenous control input. In contrast, optimal control energy first rapidly decreased and then slightly increased with larger time horizons (SFig. 1A). The increasing amount of optimal control energy might be required to additionally constrain the distance of traversed brain states over longer time horizons. In general, we found a high consistency between the metrics across a wide range of examined time horizons. However, smaller time horizons demonstrated a different control regime in which average controllability (individual level: rmin = −0.56, p = 0.09; regional level (rmin = 0.88, p = 7 × 10−78) and minimum control energy (individual level: rmin = 0.28, p = 0.44; regional level (rmin = −0.71, p = 6 × 10−37) were partly anti- correlated with the corresponding metrics in larger time horizons. In sum, short time horizons induced an alternative control regime in average controllability and minimum control energy compared to longer time horizons. Impact of normalization 4.3 The normalization step represents another modeling decision that is related to time. For mathematical reasons, we often assume the neural dynamics to diminish and stabilize over time. Neurobiological con- siderations determine the degree of normalization; that is, how fast or slow we assume the neural system to stabilize. To investigate the effect of normalization on controllability metrics and control energies, we calculated average controllability, modal controllability, and minimum control energy for different choices of the normalization parameter c. At both individual and regional levels, we first observed that with increasing c, average controllability decreased whereas modal controllability and minimum control energy increased (SFig. 2). Next, we investigated the consistency of the metrics across different manners of normalization by quantifying the Pearson correlation between metrics for two choices of the normaliza- tion parameter c, separately summarized across brain regions (Fig. 3A) and individuals (SFig. 3A). In both cases, we observed two different control regimes depending on small (c = 0.1 to c = 102; Fig. 3A) and large (c = 104 to c = 106; Fig. 3C) normalization parameters. Within each regime, the results were highly consistent independent of the normalization parameter c. Between both regimes, however, the con- sistency in average controllability (individual level: rmin = −0.19, p = 0.61; regional level: rmin = 0.86, p = 1× 10−69;), modal controllability (individual level: rmin = 0.29, p = 0.41; regional level: rmin = 0.99, p = 7× 10−320), and minimum control energy (individual level: rmin = 0.87, p = 2× 10−3; regional level: rmin = 0.81, p = 6 × 10−56) was reduced. Hypothesizing that this alternative control regime might be due to a faster stabilization of the system, we quantified the Spearman correlation between c and the decay rate of the slowest mode. We indeed found that an increase of the normalization parameter led to a faster stabilization of the system (r = −1.0, p = 0). Taken together, a faster stabilization of the system 13 introduced an alternative control regime that particularly affected controllability metrics. Figure 3: Different control regimes depending on normalization. Consistency of (Left) average controllability, (Middle) modal controllability, and (Right) minimum control energy for different choices of the normalization parameter c. (A) Heat maps depict correlation matrices of different normalization parameters. Each heat map entry corresponds to the Pearson correlation between a metric calculated using one normalization parameter and the same metric calculated using a second normalization parameter. Pearson correlation of individual metrics summarized across brain regions. Normalizing the system such that it stabilizes faster introduces a different control regime. (B) Regional controllability metrics and control energy projected onto the brain surface using a normalization parameter c = 1. (C) Analogously, the same metrics plotted onto the brain surface using a normalization parameter c = 105. Together, panels (B) and (C) illustrate the fact that distinct choices for the normalization parameter can induce distinct control regimes. Note: metric values are ranked for visualization purposes only. 14 4.4 Impact of control set size In the study of the effects of brain stimulation on brain activity, we can also ask how many and which brain regions we should control in order to drive the system to a, for instance healthy, state. More concretely, we could compare the effects of targeting a specific neural circuit to the effects of whole-brain stimulation. This motivates the examination of a final modeling choice: the number of controlled brain regions. To probe the effect of control set size on minimum control energy, we began by generating ran- dom control sets of varying number of active brain regions ranging from single-node to full-brain control. We then proceeded by testing the impact on minimum control energy and the numerical error in six brain state transitions: from activation of the default mode to activation of six canonical cognitive systems as defined by Yeo et al. [72]. Importantly, the numerical error was reasonably small (< 1 × 10−6) when we controlled at least 28.3% to 29.6% of brain regions (NV IS = 66, NSOM = 67, NDOR = 67, NV EN = 68, NLIM = 69, NF P C = 68), increasing our confidence in the results. We observed that minimum control energy and the numerical error decreased exponentially with increasing control set size (SFig. 4A). In- tuitively, the control of a larger number of brain regions required less control energy. The exponential relationship between control energy and control node set can also be mathematically derived [11]. Next, we were interested in how control and state trajectories differ in partial- compared to full-brain control sets. We calculated the minimum control energy trajectory and the distance between the state trajectory and the target state for the same six state transitions controlling all versus randomly drawn sets of 150 brain regions. In full-brain control, we observed an exponential increase in energy (Fig. 4A) and an approximately linear decrease in the distance between current and target state (Fig. 4B) across the control horizon. When we controlled only a part of the brain, control and state trajectories differed considerably. For instance, instead of taking the direct route through the state space, the system tra- versed more distant states before it reached the target state. Theoretical work has indeed shown that such non-local trajectories generally emerge if only a subset of nodes is controlled [77]. Finally, we wished to study the effect of distance between initial and target state on minimum con- trol energy. Because the set of state transitions that we studied lacked sufficient variability in these distances, we additionally simulated trajectories from a zero-activity initial state to random target states with a varying size of brain regions activated. We found a monotonic increase of minimum control energy with increasing distance between initial and target states (Fig. 4C). When employing a partial control set, a subset of the random state transitions required massive amounts of control energy. A further exploration revealed that these hardly controllable state transitions involved an activation of two weakly connected limbic regions that were not part of the random control set. Similarly, the six state transitions likely required less control on average because the activation of densely connected cognitive systems is an easier control task than the activation of randomly chosen regions in target states of equal distance. The findings in optimal control energy were highly similar, even if the exact control and state trajectories were different (SFig. 5). Overall, state and control trajectories differed substantially depending on which brain regions were allowed to receive energy input. 4.5 Relation of metrics How easily a brain network can be steered to different states and the amount of energy required to achieve a specific state transition could prove to be informative markers for pathology in or injury to the central nervous system. The selection of the metric put to the test primarily depends on the specific research question. To elucidate the empirical associations between controllability metrics and control energies, 15 Figure 4: Minimum control energy in full and partial control sets. Minimum control energy for state transitions from the default mode system to six target cognitive systems (colored lines and triangles). (Top) Results from simulations using a whole-brain control set including all 233 brain regions. (Bottom) Results from simulations using a control set consisting of random subsets of 150 brain regions. (A) Minimum control energy across the control trajectory differs quantitatively between full and partial brain control. (B) Euclidean distance between the current state and the target state across the control trajectory differs between full and partial brain control. (C) Minimum control energy increases with larger Euclidean distance between initial and target states. Blue dots depict state transitions from a zero-activity brain state to states comprised of a varying number of randomly activated brain regions. All values are averaged across participants. Lines and ribbons represent the best fit to the data and the 95% confidence interval, respectively. Abbreviations: x0 = initial state, xT = target state, x(t) = state at time t, VIS = visual, SOM = somatomotor, DOR = dorsal attention, VEN = ventral attention, LIM = limbic, and FPC = frontoparietal control. we measured the Pearson correlation between each pair of metrics, summarized separately across brain regions (Fig. 5A) and individuals (Fig. 5B and Fig. 5C). In line with previous research [31], average and modal controllability showed a positive association on the individual level (r = 0.5, p = 0.14), but asignifi- cant large negative association on the regional level (r = −0.87, p = 1×10−74). In the six state transitions we studied, we found a significant large positive correlation between minimum and optimal control energy both on the individual level (r = 0.99, p = 1× 10−7) and on the regional level (r = 0.97, p = 1× 10−136). Consistent with the mathematical constraints of their definitions, minimum control energy was signifi- cantly lower than optimal control energy on both the individual (meanmin = 43.8, meanopt = 56.0, V = 0, p = 2 × 10−3) and regional (meanmin = 0.19, meanopt = 0.24, V = 2, p = 6 × 10−40) levels. Finally, we investigated how controllability metrics and control energies are related. On both levels of examina- tion, average controllability demonstrated small negative correlations with control energies (ranging from 16 r = −0.29, p = 6 × 10−6 to r = −0.03, p = 0.93), whereas modal controllability showed slightly larger, positive associations with control energies (ranging from r = 0.23, p = 3 × 10−4 to r = 0.45, p = 0.19). The size of the correlations estimated with empirical data supported the scarcity of a clear theoretical link between the concepts. Yet, the directions of the effects was consistent with the general notion that high average controllability implies low control energy, whereas modal controllability is linked to higher control energies. Figure 5: Relation between controllability metrics and control energies. (A) The correlation matrix of individual average controllability, modal controllability, minimum control energy, and optimal control energy, summarized across brain regions. Pearson correlation coefficients (upper matrix triangle) additionally encoded via color and size of circles (lower matrix triangle). (B) Analogously, the correlation matrix of regional controllability metrics and control energies averaged across participants. (C) Regional controllability metrics and control energies projected onto the brain surface. Note that the metric values are ranked for visualization purposes only. Collectively, these panels illustrate the negative association between regional average and modal controllability, and the high consistency between minimum and optimal control energy. Abbreviations: AVG = average controllability, MOD = modal controllability, MIN = minimum control energy, and OPT = optimal control energy. Asterisks indicate significance on a Bonferroni-corrected α-level of 0.05 (*), 0.01 (**), and 0.001 (***). 4.6 Structural connectivity measures After systematically examining the impact of diverse modeling choices, we wished to provide several, potentially useful extensions of the theoretical framework. We begin with a consideration of the archi- tecture of the brain which represents the core of network control theory. Thus, it is particularly relevant how we define the inter-connections between brain regions. Typically used DTI data do not take into 17 account the fact that the signal can theoretically diffuse via physical contact between two brain regions. To evaluate the consequences of different forms of the adjacency matrix reflecting different modes of signal propagation in the brain, we additionally built structural connectivity networks based on the amount of shared neighborhood between two brain regions. Then, we calculated controllability metrics and control energies for the two alternative measures of structural connectivity, their combination, and their bina- rized versions (Fig. 6). We first examined the similarity in controllability of structural networks based on diffusion imaging (A) and based on spatial adjacency (S). Between A and S, we found small- to medium-sized Pearson correlations in average controllability (individual level: r = 0.02, p = 0.95; regional level: r = −0.01, p = 0.92), modal controllability (individual level: r = −0.15, p = 0.67; regional level: r = 0.41, p = 10 × 10−11), and minimum control energy (individual level: r = 0.36, p = 0.31; regional level: r = 0.64, p = 2× 10−16). Thus, the two measures of structural connectivity provide complementary information. Next, we quantified the effect of binarization, matrix type (A vs. S), and their combination AS, on controllability metrics and control energy. Repeated measures ANOVAs revealed significant main effects of matrix type and binarization on a Bonferroni-corrected level of α = 0.01 except for the effect of matrix type on average controllability (individual level: F = 5.98, p = 5 × 10−3; regional level F = 1.30, p = 0.27). To ensure that these results were not exclusively due to different edge weight distributions, we veri- fied these results using S and AS based on the same edge weight distribution as A. When we examined the effects in more detail, we observed that the binarization reduced the absolute values and variance of average controllability on both the regional and individual level, whereas modal controllability displayed a reverse effect. This pattern of results is in line with findings that less connected brain regions exhibit lower average controllability but higher modal controllability [36]. Similarly, individual minimum control energy was increased for binary matrices compared to fully weighted matrices; this result is consistent with previous evidence demonstrating that control nodes with more homogeneous edge weights require larger control energy [30]. Overall, the binarization of the structural connectivity matrix substantially reduced the variance of controllability metrics but not minimum control energy, suggesting that the edge weights carry valuable information especially for controllability metrics. 4.7 Persistent and transient modal controllability Many neuroscientific endeavors focus on the speed of neural dynamics. Network control theory allows us to explicitly study whether a brain region is capable of controlling fast and slowly changing activity modes by means of transient and persistent modal controllability. However, there is no clear definition of which activity modes are considered as fast or slow. Thus, we wished to further inspect how the definition of fast and slow temporal dynamics affects transient and persistent modal controllability. We began with the calculation of both metrics across various thresholds for determining which modes were considered to be transient versus persistent. First, we observed that with increasing threshold the magnitude of both transient and persistent modal controllability increased because the number of summed modes was expanded. As expected, we further noted that transient and persistent modal controllability based on a threshold of 0.5 summed up to modal controllability. The initially positive Pearson correlation between transient and persistent modal controllability of brain regions reduced and turned into a negative associ- ation with increasing thresholds (r0.1 = 0.82, p0.1 = 7 × 10−59; r0.2 = 0.78, p0.2 = 7 × 10−49; r0.3 = 0.65, p0.3 = 1 × 10−29; r0.4 = −0.20, p0.4 = 3 × 10−3; r0.5 = −0.99, p0.5 = 4 × 10−187) (Fig. 7A). Notably, for small thresholds such as 0.1, a subset of brain regions was found to be capable of controlling both fast and slow temporal dynamics (Fig. 7B). While controlling for the size of each cognitive system, we 18 Figure 6: Structural connectivity measures. (A) Average controllability, (B) modal controllability, and (C) minimum control energy for different measures of structural connectivity. The network encoded in A is based on streamline counts between two brain regions from diffusion imaging. The network encoded in S is based on the extent of spatial adjacency between two brain regions from T1-weighted images. The network encoded in AS is an average of A and S. Additionally, we consider binary versions of the three networks, and refer to them as bA, bS, and bAS, respectively. (Top) Box plots depict individual controllability metrics and control energy summarized across brain regions. Diamonds represent individuals. (Bottom) Violin plots depict regional metrics averaged across participants. Collectively, these panels illustrate the fact that the two structural connectivity measures provide complementary information that is retained by their combination. found that these brain regions belonged primarily to the subcortex (36%) and VIS (22%) systems, but also VEN (12%), DOR (9%), SOM (8%), DM (8%), and FPC (5%) systems. For large thresholds such as 0.5, brain regions seem to be either able to control fast dynamics (39% SC, 14% DOR, 12% VIS, 11% DM, 8% FPC, 6% SOM, 5% VEN, and 5% LIM systems) or slow dynamics (31% FPC, 26% subcortex, 22% SOM, 10% DOR, 7% DM, and 3% VIS systems), but not both. To explore this ambiguous relationship in more detail, we disentangled the overlapping thresholds by considering the unscaled controllability matrix V , and then by summarizing the modes into 10 intervals versus 2 intervals (Fig. 7C, top versus bottom). Interestingly, this investigation into the controllability of separate mode intervals also supported the notion that similar brain regions were capable of controlling fast and slow dynamics in the strict definition of these control tasks (10 intervals) but not in the broader definition of these control tasks (2 intervals). Importantly, we note that these results do not extend to discrete-time systems because the definition of modes that are considered as fast versus slow differs substantially between time systems. Overall, the ability of a brain region to control fast and slow modes largely depended on the definition of the control tasks. 19 Figure 7: Impact of threshold on persistent and transient modal controllability. Regional controllability of fast and slow modes for two exemplary thresholds. (Top) Persistent and transient modal controllability defined as a brain region's ability to control the 10% slowest and 10% fastest modes, respectively. (Bottom) Analogously, persistent and transient modal controllability based on a threshold of 50% of the modes. (A) Scatter plots show the relationship between transient and persistent modal controllability of brain regions averaged across participants. (B) Transient and persistent modal controllability projected onto the brain surface. Note that metric values are ranked for visualization purposes only. (C) Heat maps depict each node's ability to control a specific interval of modes, ranging from the fastest (1) to the slowest (10 and 2 respectively) modes. For this purpose, we summarized the unscaled controllability matrix V into 10 and 2 intervals respectively. When we aggregated the modes into 10 intervals, similar brain regions were capable of controlling both the slowest and fastest group of modes. When we, however, aggregated the modes into 2 intervals, brain regions were able to control either fast or slow modes. Thus, the ability of a brain region to control fast and slow modes depended on the definition of the specific control task. Abbreviations: ctrb = controllability, VIS = visual, SOM = somatomotor, DOR = dorsal attention, VEN = ventral attention, LIM = limbic, FPC = frontoparietal control, DM = default mode network, and SC = subcortical. 4.8 Complexity of energy landscape Finally, we sought to extend the types of research question we can address with the set of currently available controllability and energy metrics. For this purpose, we developed and validated a complemen- tary metric that measures the heterogeneity of all possible minimum control energy trajectories. The complexity of the energy landscape allows us to quantify the similarity or dissimilarity of all possible state transitions in respect to their required amount of control energy. Based on the variability of the eigenvalues of the controllability Gramian, we quantified the complexity of the minimum control energy landscape in each individual. Probing the consistency of the complexity of the energy landscape across time systems, we observed a large positive Pearson correlation between discrete- and continuous-time systems (r = 0.87, p = 1 × 10−3). We further examined the complementarity of the complexity of the energy landscape by calculating the Pearson correlation between the complexity measure and the other 20 established control metrics defined earlier. We found a small negative association between complexity and average controllability (r = −0.15, p = 0.68), a large negative association with modal controllability (r = −0.67, p = 0.04), and a medium negative association with minimum control energy (r = −0.40, p = 0.26). Next, we validated the complexity of the energy landscape of the brain against three null models, preserving either the strength distribution or the spatial embedding, or both. Brain networks showed a significantly lower complexity of the energy landscape than the topological null model (W = 65, p = 8 × 10−8), the spatial null model (W = 0, p = 5 × 10−8), and the combined null model (W = 2498, p = 6× 10−3), as quantified by a Wilcoxon test (Fig. 8). Interestingly, the combination of topological and spatial characteristics seemed to partially explain the brain's higher homogeneity of the energy landscape. We found consistent evidence in discrete-time systems (SFig. 7). Overall, the complexity of the energy landscape of the brain was complementary to other controllability metrics and low compared to several null models. Figure 8: Complexity of the energy landscape of the human brain. Heterogeneity of the minimum control energy landscape of individual participants (dark blue diamonds) as compared to three null models preserving different characteris- tics of brain networks. The complexity of the energy landscape was quantified by the variability of the eigenvalue distribution of the controllability Gramian. Null model distributions (box plots) were estimated by randomly rewiring each brain net- work 100 times. Spatial null models (blue box plots) preserved the relationship between edge weight and Euclidean distance. Topological null models (yellow box plots) preserved degree and strength distributions. Combined null models (green box plots) preserved both strength distribution and spatial embedding. Dashed lines indicate complexity of the energy landscape of brain networks and null models averaged across individuals. The combination of topological and spatial characteristics partially explains the homogeneous energy landscape of the brain. 5 Discussion Network control theory is an emerging field in neuroscience that has the potential to yield promising insights into structure-function relationships in health and disease. Here, we provided an overview of the theoretical framework by illustrating the underlying model of neural dynamics and commonly stud- ied controllability concepts. Based on the structural brain networks from ultra high-resolution diffusion imaging data (730 diffusion directions) of 10 healthy adults, we calculated average and modal control- lability as well as minimum and optimal control energy. We then systematically probed the impact of 21 different modeling choices, specifically the choice of time system, time horizon, normalization, and size of the control set, on these metrics. We further suggested potentially useful model extensions such as an alternative measure of structural connectivity accounting for propagation of signals through gray matter to abutting regions, and a complementary metric quantifying the complexity of the energy landscape of brain networks. Specific modeling recommendations. Based on our systematic examination of different modeling choices, we derived several specific recommendations. First, we observed a generally high consistency be- tween the behavior of discrete- and continuous-time systems, which depended on the metric, observation level, and time horizon. Classifying the neural dynamics under study as clearly discrete- or continuous- time is often challenging. Unless an investigator has a clear justification for choosing one time system over another, we recommend to verify the obtained results in the alternative time-system to allow for a better generality of the findings and inferences drawn therefrom. Second, we demonstrated that short time horizons led to an alternative time system compared to longer time horizons. The arbitrary units of the time scale further challenge the decision of which time horizon to choose. If there exists no concrete justification for the choice of time horizon, we recommend to validate the obtained findings using several different time horizons. Third, we found that a fast stabilization of the system induced a substantially different control scenario. Again if there are no concrete neurobiological variables that can be used to constrain one's choice, we suggest that a slow stabilization could be a plausible representation of most neural dynamics, allowing for a broader range of dynamics. Since the influence of the normalization pa- rameter c depends on the largest eigenvalues, the same c can have different stabilization effects in different brain networks. To ensure consistency across studies, we suggest to make c dependent on λ(A)max, for instance by c = 0.01·λ(A)max. Finally, we observed that the composition and size of the control region set substantially influenced state and control trajectories. The decision critically depends on the individ- ual research question and hence, should be well informed by theoretical or practical considerations. From a methodological perspective, it is important to control a sufficiently large number of brain regions to robustly estimate control energies. In sum, these recommendations could guide more informed modeling choices in future applications of network control theory to pressing questions in cognitive, developmental, and clinical neuroscience. The role of time in network controllability. In our examination of different modeling choices, we found that both a short time horizon and a fast stabilization of the system induced an alternative control regime. We suggest a common mechanism underlying both time-related observations. Whereas the injected control input has time to diffuse along inter-connections between brain regions over longer time horizons, it might be possible that this diffusion process is constrained over short time horizons. Instead, a different control regime could come into effect in which the injected input primarily controls each brain region independently rather than capitalizing on their interconnections. This finding suggests that time might play a more important role in the controllability of structural brain networks than is commonly assumed. Thus, it could be interesting to further investigate the factor of time, for instance by linking control to real-time measures of brain function [18, 58]. Another potentially fruitful venture could be to determine optimal control horizons by capitalizing on the natural dynamics of the system or by changing inter-connections in more advanced dynamic models [78]. Such methods emphasizing the role of time could help to develop minimal clinical interventions such as neuromodulation [79], which is immediately relevant for the control of seizures in epilepsy [80 -- 84]. The temporal nature of control is also potentially relevant for further refining brain-machine interfaces [85, 86]. 22 Future directions for proposed model extensions. Moreover, the present work provides several potentially useful extensions of network control theory. We first developed and validated a complemen- tary measure of structural connectivity motivated by the fact that brain networks based on diffusion imaging data disregard the potential for neural signals to diffuse between spatially adjacent brain re- gions. We demonstrated that this alternative structural connectivity measure based on the amount of shared neighborhood between two brain regions was complementary to the tractography version. We fur- ther showed that their combination introduced more inter-individual variability in controllability metrics, motivating future efforts to employ this approach in studies of individual differences. An important next step is to test whether structural brain networks based on both diffusion imaging and spatial adjacency outperform networks purely based on diffusion imaging data by better accounting for the observed neural dynamics [45, 46]. Additionally, we examined the ability of the brain to control slow and fast dynamics. We found that the capability of a brain region to control different fast modes depended on the specific definition of the control task and was not consistent between time-systems. Neuroscientists interested in the speed of neural changes such as different frequency bands [87, 88] should be careful in justifying their choice of time system and the threshold which defines slow versus fast modes. Lastly, we wished to extend the existing set of controllability metrics. For this purpose, we developed and validated a new metric that quantifies the complexity of the energy landscape of a given brain network. In other words, the metric measures how heterogeneous all possible state transitions are in the control energy that they require. We showed that the brain exhibited a more homogeneous energy landscape compared to two different null models. We found that both the brain networks' strength distribution and spatial embedding partially explained this observation, which is in line with previous findings connecting local and global network characteristics to network controllability [30, 31, 36]. The requirement of a sim- ilar amount of energy to enable diverse state transitions implies that brain architecture supports diverse transitions, which in turn could explain the complex functional dynamics consistently observed in neural systems. A crucial next step is to test the practical utility of this new metric by linking it to development, cognition, and psychiatric disorders. Taken together, the proposed model extensions hopefully stimulate and enrich future research. Expanding horizons of network control theory. New developments in network control theory are constantly expanding the horizons of research questions that can be tackled with the associated tools. Many (but perhaps not all) of these developments could be helpful in the study of the mind and brain. Efforts have recently revealed a relation between controllability and symmetry [89 -- 91], which could prove useful in determining the impact of bilateral and other symmetries on neural dynamics. The field has begun considering multiobjective functions, tradeoffs, and constraints in control [92, 93], in addition to probing a system's potential for control via local topological information [94]. As the field of neu- roscience moves more concertedly towards multimodal approaches, efforts in the control of multilayer networks could prove particularly useful [95], as could methods for detecting control nodes across scales [96 -- 98]. For some questions, advances in the control of nonlinear systems could prove effective [99 -- 101], including applications of Ising models [102, 103] and considerations related to the dynamics of neural mass models [104]. Finally, moving beyond network controllability, recent work expanding system identi- fication methods to identify specific form of nonlinear dynamics present in brain is particularly promising [105 -- 107]. 23 Methodological considerations. Several methodological aspects could potentially constrain the in- terpretability of our results. First, we capitalized on high-resolution diffusion weighted imaging data for the construction of structural connectivity networks. Associated tractography algorithms are still limited in their capacity to reliably track fiber bundles, particularly long-range connections [108, 109], in terms of their origin, exact direction, and intersection [110]. Nevertheless, diffusion weighted imaging serves as the state-of-the art method to study white matter architecture in humans and therefore, tractography algorithms are continuously being refined [111]. Second, our dynamic model of neural processes relied on several simplifying assumptions including linearity and time-invariance. However, such basic models often provide a good starting-point to approximate higher-order dynamics [45, 46] and can subsequently be adapted to contain more complex features such as non-linearity [101, 112] and time-dependence [78]. Third, it was beyond the scope of this paper to examine the impact of modeling choices in all of their theoretically possible combinations. Instead, we systematically varied one modeling choice at a time while keeping all other choices constant. Thus, the obtained results might not automatically generalize to left-out choices, for example in the presence of higher order interactions. A further limitation of our study is the investigation of control energies in a restricted set of six state transitions. In addition, the initial and target brains states were constructed in a controlled, yet unnatural way by capitalizing on the artificial activation of brain regions belonging to the same cognitive system. For greater ecological validity, future studies could instead rely on real brain states as measured by functional neuroimaging [18, 58]. Lastly, we wish to point out that the high consistency between minimum and optimal control energy could also be due to different scales of distance and energy costs. To avoid such effects, future efforts could develop an optimal energy algorithm that balances both constraints equally independent of their scale. Conclusions. Our systematic overview of network control theory and possible modeling choices aimed to facilitate a deeper understanding and better evaluation of network control theory applications in neuro- science. Future work can potentially benefit from our specific recommendations and the proposed model extensions. Overall, this work hopefully inspires the neuroscience community to fully exploit the poten- tial of network control theory on multiple spatial scales ranging from single neurons to brain regions. Ultimately, such endeavors could advance our understanding of how the architecture of the brain gives rise to complex neural dynamics. 6 Conflict of Interest The authors declare no competing interests. 7 Acknowledgements TMK was supported by the International Research Training Group (IRTG 2150) of the German Research Foundation and by a full doctoral scholarship of the German National Academic foundation. We gratefully acknowledge Eli J. Cornblath, Xiaosong He, Urs Braun, and Lorenzo Caciagli for useful discussions. DSB acknowledges support from the John D. and Catherine T. MacArthur Foundation, the Alfred P. Sloan Foundation, the ISI Foundation, the Paul Allen Foundation, the Army Research Laboratory (W911NF-10- 2-0022), the Army Research Office (Bassett-W911NF-14-1-0679, DCIST- W911NF-17-2-0181, Grafton- W911NF-16-1-0474), the Office of Naval Research, the National Institute of Mental Health (2-R01-DC- 24 009209-11, R01-MH112847, R01-MH107235, R21-M MH-106799), the National Institute of Child Health and Human Development (1R01HD086888-01), National Institute of Neurological Disorders and Stroke (R01 NS099348), and the National Science Foundation (BCS-1441502, BCS-1430087, NSF PHY-1554488 and BCS-1631550). The content is solely the responsibility of the authors and does not necessarily represent the official views of any of the funding agencies. 8 Author Contributions DSB, JS, and TMK designed the study. AEK preprocessed the data. JZK contributed analytic solutions. TMK and JZK wrote code. TMK analyzed the data and wrote the manuscript. All authors edited the manuscript approved the final version. 25 References [1] G Deco, V K Jirsa, and A R McIntosh. Emerging concepts for the dynamical organization of resting-state activity in the brain. Nature Reviews Neuroscience, 12(1):43, 2011. [2] G Deco and M L Kringelbach. Great expectations: using whole-brain computational connectomics for understanding neuropsychiatric disorders. Neuron, 84(5):892 -- 905, 2014. [3] A M Hermundstad, K S Brown, D S Bassett, and J M Carlson. Learning, memory, and the role of neural network architecture. PLoS computational biology, 7(6):e1002063, 2011. [4] A M Hermundstad, D S Bassett, K S Brown, E M Aminoff, D Clewett, S Freeman, A Frithsen, A Johnson, C M Tipper, M B Miller, et al. Structural foundations of resting-state and task-based functional connectivity in the human brain. Proceedings of the National Academy of Sciences, 110 (15):6169 -- 6174, 2013. [5] A M Hermundstad, K S Brown, D S Bassett, E M Aminoff, A Frithsen, A Johnson, C M Tipper, M B Miller, S T Grafton, and J M Carlson. Structurally-constrained relationships between cognitive states in the human brain. PLoS computational biology, 10(5):e1003591, 2014. [6] K Rajan, C D Harvey, and D W Tank. Recurrent network models of sequence generation and memory. Neuron, 90(1):128 -- 142, 2016. [7] N Levy, D Horn, I Meilijson, and E Ruppin. Distributed synchrony in a hebbian cell assembly of spiking neurons. Advances in Neural Information Processing Systems, 11, 1999. [8] I R Fiete, W Senn, C ZH Wang, and R HR Hahnloser. Spike-time-dependent plasticity and het- erosynaptic competition organize networks to produce long scale-free sequences of neural activity. Neuron, 65(4):563 -- 576, 2010. [9] T Kailath. Linear systems, volume 156. Prentice-Hall Englewood Cliffs, NJ, 1980. [10] Y-Y Liu, J-J Slotine, and journal=nature volume=473 number=7346 pages=167 year=2011 pub- lisher=Nature Publishing Group Barab´asi, A-L. Controllability of complex networks. [11] F Pasqualetti, S Zampieri, and F Bullo. Controllability metrics, limitations and algorithms for complex networks. IEEE Transactions on Control of Network Systems, 1(1):40 -- 52, 2014. [12] E Bullmore and O Sporns. Complex brain networks: graph theoretical analysis of structural and functional systems. Nature reviews neuroscience, 10(3):186, 2009. [13] D S Bassett, P Zurn, and J I Gold. On the nature and use of models in network neuroscience. Nat Rev Neurosci, 19(9):566 -- 578, 2018. [14] K V Shenoy, M T Kaufman, M Sahani, and M M Churchland. A dynamical systems view of motor preparation: implications for neural prosthetic system design. In Progress in brain research, volume 192, pages 33 -- 58. Elsevier, 2011. [15] W J Freeman. Characterization of state transitions in spatially distributed, chaotic, nonlinear, dynamical systems in cerebral cortex. Integrative Physiological and Behavioral Science, 29(3):294 -- 306, 1994. 26 [16] R E Kalman, Y-C Ho, and K S Narendra. Controllability of linear dynamical systems, contributions to differential equations 1 (1963), 189 -- 213. MR, 155070(27):5012. [17] S F Muldoon, F Pasqualetti, S Gu, M Cieslak, S T Grafton, J M Vettel, and D S Bassett. Stimulation-based control of dynamic brain networks. PLoS Comput Biol, 12(9):e1005076, 2016. [18] J Stiso, A N Khambhati, T Menara, A E Kahn, J M Stein, S R Das, R Gorniak, J Tracy, B Litt, K A Davis, et al. White matter network architecture guides direct electrical stimulation through optimal state transitions. arXiv preprint arXiv:1805.01260, 2018. [19] E A Solomon, J E Kragel, R Gross, B Lega, M R Sperling, G Worrell, S A Sheth, K A Zaghloul, B C Jobst, J M Stein, S Das, R Gorniak, C S Inman, S Seger, D S Rizzuto, and M J Kahana. Medial temporal lobe functional connectivity predicts stimulation-induced theta power. Nat Commun, 9 (1):4437, 2018. [20] J D Medaglia, D Y Harvey, N White, A Kelkar, J Zimmerman, D S Bassett, and R H Hamilton. Network controllability in the inferior frontal gyrus relates to controlled language variability and susceptibility to tms. Journal of Neuroscience, pages 0092 -- 17, 2018. [21] A N Khambhati, A E Kahn, J Costantini, Y Ezzyat, E A Solomon, R E Gross, B C Jobst, S A Sheth, K A Zaghloul, G Worrell, S Seger, B C Lega, S Weiss, M R Sperling, R Gorniak, S R Das, J M Stein, D S Rizzuto, M J Kahana, T H Lucas, K A Davis, J I Tracy, and D S Bassett. Functional control of electrophysiological network architecture using direct neurostimulation in humans. Network Neuroscience, Epub Ahead of Pring, 2019. [22] Z Cui, J Stiso, G L Baum, J Z Kim, D R Roalf, R F Betzel, Shi Gu, Z Lu, C H Xia, R Ciric, et al. Optimization of energy state transition trajectory supports the development of executive function during youth. bioRxiv, page 424929, 2018. [23] S F Muldoon, J Costantini, W R S Webber, R Lesser, and D S Bassett. Locally stable brain states predict suppression of epileptic activity by enhanced cognitive effort. Neuroimage Clin, 18:599 -- 607, 2018. [24] E J Cornblath, E Tang, G L Baum, T M Moore, A Adebimpe, D R Roalf, R C Gur, R E Gur, F Pasqualetti, T D Satterthwaite, et al. Sex differences in network controllability as a predictor of executive function in youth. NeuroImage, 188:122 -- 134, 2019. [25] R P Dum, D J Levinthal, and P L Strick. Motor, cognitive, and affective areas of the cerebral cortex influence the adrenal medulla. Proc Natl Acad Sci U S A, 113(35):9922 -- 9927, 2016. [26] M S Goyal, S Venkatesh, J Milbrandt, J I Gordon, and M E Raichle. Feeding the brain and nurturing the mind: Linking nutrition and the gut microbiota to brain development. Proc Natl Acad Sci U S A, 112(46):14105 -- 14112, 2015. [27] G Yan, P E Vrtes, E K Towlson, Y L Chew, D S Walker, W R Schafer, and A L Barabsi. Network control principles predict neuron function in the Caenorhabditis elegans connectome. Nature, 550 (7677):519 -- 523, 2017. [28] E K Towlson, P E V´ertes, G Yan, Y L Chew, D S Walker, W R Schafer, and A Barab´asi. Caenorhab- ditis elegans and the network control frameworkfaqs. Philosophical Transactions of the Royal Society B: Biological Sciences, 373(1758):20170372, 2018. 27 [29] L Wiles, S Gu, F Pasqualetti, B Parvesse, D Gabrieli, D S Bassett, and D F Meaney. Autaptic connections shift network excitability and bursting. Sci Rep, 7:44006, 2017. [30] J Z Kim, J M Soffer, A E Kahn, J M Vettel, F Pasqualetti, and D S Bassett. Role of graph architecture in controlling dynamical networks with applications to neural systems. Nature physics, 14(1):91, 2018. [31] S Gu, F Pasqualetti, M Cieslak, Q K Telesford, B Y Alfred, A E Kahn, J D Medaglia, J M Vettel, M B Miller, S T Grafton, et al. Controllability of structural brain networks. Nature communications, 6:8414, 2015. [32] J M Shine, M Breakspear, P T Bell, K A Ehgoetz Martens, R Shine, O Koyejo, O Sporns, and R A Poldrack. Human cognition involves the dynamic integration of neural activity and neuromodula- tory systems. Nat Neurosci, 22(2):289 -- 296, 2019. [33] J Jeganathan, A Perry, D S Bassett, G Roberts, P B Mitchell, and M Breakspear. Fronto-limbic dysconnectivity leads to impaired brain network controllability in young people with bipolar disorder and those at high genetic risk. Neuroimage Clin, 19:71 -- 81, 2018. [34] B C Bernhardt, F Fadaie, M Liu, B Caldairou, S Gu, E Jefferies, J Smallwood, D S Bassett, A Bernasconi, and N Bernasconi. Temporal lobe epilepsy: Hippocampal pathology modulates connectome topology and controllability. Neurology, 92(19):e2209 -- e2220, 2019. [35] S Gu, R F Betzel, M G Mattar, M Cieslak, P R Delio, S T Grafton, F Pasqualetti, and D S Bassett. Optimal trajectories of brain state transitions. Neuroimage, 148:305 -- 317, 2017. [36] R F Betzel, S Gu, J D Medaglia, F Pasqualetti, and D S Bassett. Optimally controlling the human connectome: the role of network topology. Scientific reports, 6:30770, 2016. [37] D S Bassett and E T Bullmore. Human brain networks in health and disease. Current opinion in neurology, 22(4):340, 2009. [38] J D Medaglia, M-E Lynall, and D S Bassett. Cognitive network neuroscience. Journal of cognitive neuroscience, 27(8):1471 -- 1491, 2015. [39] M P Van den Heuvel, E T Bullmore, and O Sporns. Comparative connectomics. Trends in cognitive sciences, 20(5):345 -- 361, 2016. [40] U Braun, A Schaefer, R F Betzel, H Tost, A Meyer-Lindenberg, and D S Bassett. From maps to multi-dimensional network mechanisms of mental disorders. Neuron, 97(1):14 -- 31, 2018. [41] E Tang, C Giusti, G L Baum, S Gu, E Pollock, A E Kahn, D R Roalf, T M Moore, K Ruparel, R C Gur, et al. Developmental increases in white matter network controllability support a growing diversity of brain dynamics. Nature Communications, 8(1):1252, 2017. [42] U Braun, A Harneit, G Pergola, T Menara, A Schaefer, R F Betzel, Z Zang, J I Schweiger, K Schwarz, J Chen, G Blasi, A Bertolino, D Durstewitz, F Pasqualetti, E Schwarz, A Meyer- Lindenberg, D S Bassett, and H Tost. Brain state stability during working memory is explained by network control theory, modulated by dopamine d1/d2 receptor function, and diminished in schizophrenia. arxiv, Epub ahead of print, 2019. 28 [43] E Tang and D S Bassett. Colloquium: Control of dynamics in brain networks. Reviews of Modern Physics, 90(3):031003, 2018. [44] J Z Kim and D S Bassett. Linear dynamics & control of brain networks. arXiv preprint arXiv:1902.03309, 2019. [45] R F Gal´an. On how network architecture determines the dominant patterns of spontaneous neural activity. PloS one, 3(5):e2148, 2008. [46] CJ Honey, O Sporns, Leila Cammoun, Xavier Gigandet, Jean-Philippe Thiran, Reto Meuli, and Patric Hagmann. Predicting human resting-state functional connectivity from structural connec- tivity. Proceedings of the National Academy of Sciences, 106(6):2035 -- 2040, 2009. [47] A May. Experience-dependent structural plasticity in the adult human brain. Trends in cognitive sciences, 15(10):475 -- 482, 2011. [48] G L Baum, R Ciric, D R Roalf, R F Betzel, T M Moore, R T Shinohara, A E Kahn, S N Vandekar, P E Rupert, M Quarmley, P A Cook, M A Elliott, K Ruparel, R E Gur, R C Gur, D S Bassett, and T D Satterthwaite. Modular segregation of structural brain networks supports the development of executive function in youth. Curr Biol, 27(11):1561 -- 1572.e8, 2017. [49] B A Brinkman, A I Weber, F Rieke, and E Shea-Brown. How do efficient coding strategies depend on origins of noise in neural circuits? PLoS Comput Biol, 12(10):e1005150, 2016. [50] W F Mlynarski and A M Hermundstad. Adaptive coding for dynamic sensory inference. Elife, e32055, 2018. [51] Y Gai, B Doiron, and J Rinzel. Slope-based stochastic resonance: how noise enables phasic neurons to encode slow signals. PLoS Comput Biol, 6(6), 2010. [52] D D Garrett, A R McIntosh, and C L Grady. Brain signal variability is parametrically modifiable. Cereb Cortex, 24(11):2931 -- 2940, 2014. [53] M Breakspear and A R McIntosh. Networks, noise and models: reconceptualizing the brain as a complex, distributed system. Neuroimage, 58(2):293 -- 295, 2011. [54] L L Gollo, C Mirasso, O Sporns, and M Breakspear. Mechanisms of zero-lag synchronization in cortical motifs. PLoS Comput Biol, 10(4):e1003548, 2014. [55] P N Taylor, Y Wang, M Goodfellow, J Dauwels, F Moeller, U Stephani, and G Baier. A computa- tional study of stimulus driven epileptic seizure abatement. PLoS One, 9(12):e114316, 2014. [56] V G Boltyanskii, R V Gamkrelidze, and L S Pontryagin. The theory of optimal processes. i. the maximum principle. Technical report, TRW SPACE TECHNOLOGY LABS LOS ANGELES CALIF, 1960. [57] J P Hespanha. Linear systems theory. Princeton university press, 2018. [58] E J Cornblath, A Ashourvan, J Z Kim, R F Betzel, R Ciric, G L Baum, X He, K Ruparel, T M Moore, R C Gur, et al. Context-dependent architecture of brain state dynamics is explained by white matter connectivity and theories of network control. arXiv preprint arXiv:1809.02849, 2018. 29 [59] D Zller, C Sandini, M Schaer, S Eliez, D S Bassett, and D Van De Ville. Structural control energy of resting-state functional brain states reveals inefficient brain dynamics in psychosis vulnerability. bioRxiv, 703561, 2019. [60] T Menara, D Bassett, and F Pasqualetti. Structural controllability of symmetric networks. IEEE Transactions on Automatic Control, 2018. [61] B Marx, D Koenig, and Didier Georges. Optimal sensor and actuator location for descriptor systems using generalized gramians and balanced realizations. In Proceedings of the 2004 American Control Conference, volume 3, pages 2729 -- 2734. IEEE, 2004. [62] H R Shaker and M Tahavori. Optimal sensor and actuator location for unstable systems. Journal of Vibration and Control, 19(12):1915 -- 1920, 2013. [63] AMA Hamdan and AH Nayfeh. Measures of modal controllability and observability for first-and second-order linear systems. Journal of guidance, control, and dynamics, 12(3):421 -- 428, 1989. [64] K Gorgolewski, C D Burns, C Madison, D Clark, Y O Halchenko, M L Waskom, and S S Ghosh. Nipype: a flexible, lightweight and extensible neuroimaging data processing framework in python. Front Neuroinform, 5, 08 2011. ISSN 1662-5196. [65] B B Avants, N J Tustison, G Song, P A Cook, A Klein, and J C Gee. A reproducible evaluation of ants similarity metric performance in brain image registration. Neuroimage, 54(3):2033 -- 2044, 2011. [66] F-C Yeh, V J Wedeen, and W-Y I Tseng. Generalized q-sampling imaging. IEEE transactions on medical imaging, 29(9):1626 -- 1635, 2010. [67] F-C Yeh, T D Verstynen, Y Wang, J C Fern´andez-Miranda, and W-Y I Tseng. Deterministic diffusion fiber tracking improved by quantitative anisotropy. PloS one, 8(11):e80713, 2013. [68] M Cieslak and ST Grafton. Local termination pattern analysis: a tool for comparing white matter morphology. Brain imaging and behavior, 8(2):292 -- 299, 2014. [69] D S Bassett, J A Brown, V Deshpande, J M Carlson, and S T Grafton. Conserved and variable architecture of human white matter connectivity. Neuroimage, 54(2):1262 -- 1279, 2011. [70] L Cammoun, X Gigandet, D Meskaldji, J P Thiran, O Sporns, K Q Do, P Maeder, R Meuli, and P Hagmann. Mapping the human connectome at multiple scales with diffusion spectrum mri. Journal of neuroscience methods, 203(2):386 -- 397, 2012. [71] A Daducci, S Gerhard, A Griffa, A Lemkaddem, L Cammoun, X Gigandet, R Meuli, P Hagmann, and J-P Thiran. The connectome mapper: an open-source processing pipeline to map connectomes with mri. PloS one, 7(12):e48121, 2012. [72] B. T. Yeo, F. M. Krienen, J. Sepulcre, M. R. Sabuncu, D. Lashkari, M. Hollinshead, J. L. Roffman, Smoller J. W., L. Zllei, J. R. Polimeni, B. Fischl, ... Liu, H., and R. L. Buckner. The organization of the human cerebral cortex estimated by intrinsic functional connectivity. Journal of neurophys- iology, 106(3):1125 -- 65, 2011. 30 [73] E Wu-Yan, R F Betzel, E Tang, S Gu, F Pasqualetti, and D S Bassett. Benchmarking measures of network controllability on canonical graph models. Journal of Nonlinear Science, pages 1 -- 39, 2018. [74] E Tang, G L Baum, D R Roalf, T D Satterthwaite, F Pasqualetti, and D S Bassett. The control of brain network dynamics across diverse scales of space and time. arXiv preprint arXiv:1901.07536, 2019. [75] M Rubinov and O Sporns. Complex network measures of brain connectivity: uses and interpreta- tions. Neuroimage, 52(3):1059 -- 1069, 2010. [76] J A Roberts, A Perry, A R Lord, G Roberts, P B Mitchell, R E Smith, F Calamante, and M Break- spear. The contribution of geometry to the human connectome. Neuroimage, 124:379 -- 393, 2016. [77] J Sun and A E Motter. Controllability transition and nonlocality in network control. Physical review letters, 110(20):208701, 2013. [78] A Li, S P Cornelius, Y-Y Liu, L Wang, and A-L Barab´asi. The fundamental advantages of temporal networks. Science, 358(6366):1042 -- 1046, 2017. [79] P Afshar, X Wei, M Lazarewicz, R Gupta, G Molnar, and T Denison. Advancing neuromodulation using a dynamic control framework. Conf Proc IEEE Eng Med Biol Soc, 2011:671 -- 674, 2011. [80] S Ching, E N Brown, and M A Kramer. Distributed control in a mean-field cortical network model: implications for seizure suppression. Phys Rev E Stat Nonlin Soft Matter Phys, 86(2 Pt 1):021920, 2012. [81] S Stanslaski, J Giftakis, P Stypulkowski, D Carlson, P Afshar, P Cong, and T Denison. Emerging technology for advancing the treatment of epilepsy using a dynamic control framework. Conf Proc IEEE Eng Med Biol Soc, 2011:753 -- 756, 2011. [82] P N Taylor, J Thomas, N Sinha, J Dauwels, M Kaiser, T Thesen, and J Ruths. Optimal control based seizure abatement using patient derived connectivity. Front Neurosci, 9:202, 2015. [83] S Olmi, S Petkoski, M Guye, F Bartolomei, and V Jirsa. Controlling seizure propagation in large- scale brain networks. PLoS Comput Biol, 15(2):e1006805, 2019. [84] D Ehrens, D Sritharan, and S V Sarma. Closed-loop control of a fragile network: application to seizure-like dynamics of an epilepsy model. Front Neurosci, 9:58, 2015. [85] S Gowda, A L Orsborn, and J M Carmena. Parameter estimation for maximizing controllability of linear brain-machine interfaces. Conf Proc IEEE Eng Med Biol Soc, 2012:1314 -- 1317, 2012. [86] J Stiso, M-C Corsi, J O Garcia, J M Vettel, F De Vico Fallani, T H Lucas, and D S Bassett. Learn- ing in brain-computer interface control evidenced by joint decomposition of brain and behavior. psyArxiv, en69j, 2019. [87] M Siegel, T H Donner, and A K Engel. Spectral fingerprints of large-scale neuronal interactions. Nature Reviews Neuroscience, 13(2):121, 2012. [88] S Baillet. Magnetoencephalography for brain electrophysiology and imaging. Nature neuroscience, 20(3):327, 2017. 31 [89] A J Whalen, S N Brennan, T D Sauer, and S J Schiff. Observability and controllability of nonlinear networks: The role of symmetry. Phys Rev X, 5:1, 2015. [90] C Zhao, W X Wang, Y Y Liu, and J J Slotine. Intrinsic dynamics induce global symmetry in network controllability. Sci Rep, 5(8422), 2015. [91] A J Whalen, S N Brennan, T D Sauer, and S J Schiff. Effects of symmetry on the structural controllability of neural networks: A perspective. Proc Am Control Conf, 2016:5785 -- 5790, 2016. [92] Y Tang, H Gao, W Du, J Lu, A V Vasilakos, and J Kurths. Robust multiobjective controllability of complex neuronal networks. IEEE/ACM Trans Comput Biol Bioinform, 13(4):778 -- 791, 2016. [93] H Keren and S Marom. Controlling neural network responsiveness: tradeoffs and constraints. Front Neuroeng, 7:11, 2014. [94] G Li, L Deng, G Xiao, P Tang, C Wen, W Hu, J Pei, L Shi, and H E Stanley. Enabling controlling complex networks with local topological information. Sci Rep, 8(1):4593, 2018. [95] G Menichetti, L Dall'Asta, and G Bianconi. Control of multilayer networks. Sci Rep, 6:20706, 2016. [96] Y Tang, H Gao, and J Kurths. Multiobjective identification of controlling areas in neuronal net- works. IEEE/ACM Trans Comput Biol Bioinform, 10(3):708 -- 720, 2013. [97] Y Tang, Z Wang, H Gao, S Swift, and J Kurths. A constrained evolutionary computation method for detecting controlling regions of cortical networks. IEEE/ACM Trans Comput Biol Bioinform, 9(6):1569 -- 1581, 2012. [98] Y Tang, H Gao, W Zou, and J Kurths. Identifying controlling nodes in neuronal networks in different scales. PLoS One, 7(7):e41375, 2012. [99] S P Cornelius, W L Kath, and A E Motter. Realistic control of network dynamics. Nat Commun, 4:1942, 2013. [100] A E Motter. Networkcontrology. Chaos, 25(9):097621, 2015. [101] J G T Zanudo, G Yang, and R Albert. Structure-based control of complex networks with nonlinear dynamics. Proceedings of the National Academy of Sciences, 114(28):7234 -- 7239, 2017. [102] C W Lynn and D D Lee. Statistical mechanics of influence maximization with thermal noise. European Physics Letters, 117, 2017. [103] S Gu, M Cieslak, B Baird, S F Muldoon, S T Grafton, F Pasqualetti, and D S Bassett. The energy landscape of neurophysiological activity implicit in brain network structure. Sci Rep, 8(1):2507, 2018. [104] X Liu, J Gao, G Wang, and Z W Chen. Controllability analysis of the neural mass model with dynamic parameters. Neural Comput, 29(2):485 -- 501, 2017. [105] C O Becker, D S Bassett, and V M Preciado. Large-scale dynamic modeling of task-fMRI signals via subspace system identification. J Neural Eng, 15(6):066016, 2018. 32 [106] A Ashourvan, S Pequito, M Bertolero, J Z Kim, D S Bassett, and B Litt. A dynamical systems framework to uncover the drivers of large-scale cortical activity. bioRxiv, 638718, 2019. [107] G Koppe, J Toutounji, P Kirsch, S Lis, and D Durstewitz. Identifying nonlinear dynamical systems via generative recurrent neural networks with applications to fMRI. arXiv, 1902:07186, 2019. [108] C Thomas, Q Y Frank, M O Irfanoglu, P Modi, K S Saleem, D A Leopold, and C Pierpaoli. Anatomical accuracy of brain connections derived from diffusion mri tractography is inherently limited. Proceedings of the National Academy of Sciences, 111(46):16574 -- 16579, 2014. [109] C Reveley, A K Seth, C Pierpaoli, A C Silva, D Yu, R C Saunders, D A Leopold, and Q Y Frank. Superficial white matter fiber systems impede detection of long-range cortical connections in diffusion mr tractography. Proceedings of the National Academy of Sciences, 112(21):E2820 -- E2828, 2015. [110] S Jbabdi and H Johansen-Berg. Tractography: where do we go from here? Brain connectivity, 1 (3):169 -- 183, 2011. [111] F Pestilli, J D Yeatman, A Rokem, K N Kay, and B A Wandell. Evaluation and statistical inference for human connectomes. Nature methods, 11(10):1058, 2014. [112] B Fiedler, A Mochizuki, G Kurosawa, and D Saito. Dynamics and control at feedback vertex sets. i: Informative and determining nodes in regulatory networks. Journal of Dynamics and Differential Equations, 25(3):563 -- 604, 2013. 33
1501.06890
1
1501
2015-01-27T20:03:50
Comment on "Human Time-Frequency Acuity Beats the Fourier Uncertainty Principle"
[ "q-bio.NC" ]
In the initial article [Phys. Rev. Lett. 110, 044301 (2013), arXiv:1208.4611] it was claimed that human hearing can beat the Fourier uncertainty principle. In this Comment, we demonstrate that the experiment designed and implemented in the original article was ill-chosen to test Fourier uncertainty in human hearing.
q-bio.NC
q-bio
Comment on "Human Time-Frequency Acuity Beats the Fourier Uncertainty Principle" National Research Council of Canada, 100 Sussex Drive, Ottawa, ON, Canada K1A 0R6 G.S. Thekkadath and Michael Spanner (Dated: October 10, 2018) In the initial article [Phys. Rev. Lett. 110, 044301 (2013), arXiv:1208.4611] it was claimed that human hearing can beat the Fourier uncertainty principle. In this Comment, we demonstrate that the experiment designed and implemented in the original article was ill-chosen to test Fourier uncertainty in human hearing. 5 1 0 2 n a J 7 2 ] . C N o i b - q [ 1 v 0 9 8 6 0 . 1 0 5 1 : v i X r a The Gabor limit [1], ∆t∆f ≥ 1 4π , (1) refers to the lower bound on the product of the standard deviations (STD) in time (∆t) and frequency (∆f ) of an audio signal. This limit is a consequence of the Fourier uncertainty principle. In their Letter, Oppenheim and Magnasco [2] claim that human hearing can surpass this limit. They design an experiment which establishes psy- chological limens, δt and δf , and show that their subjects can discriminate signals that beat a limen-based uncer- tainty δtδf ≥ 1 4π . (2) The frequency and time limens used by the authors re- late to the accuracy with which human participants can distinguish small frequency and time shifts present in a sequence of three test pulses. The sequence in question is referred to as "Task 5" in the paper. It is our view that their experiment is ill-chosen to test Fourier uncertainty. Firstly, the ∆t and ∆f that appear in the Gabor limit must be the STD of time and frequency evaluated over the whole test signal. This point is made clear in the derivation of the uncertainty principle that can be found in the book of Cohen [3] (Sections 3.2 and 3.3). The li- mens used by the authors, however, are simply ad hoc parameters that relate to the STD of statistical errors made by the human participants when tasked with esti- mating frequency and timing shifts in the test signal, and are unrelated to the STD of time and frequency evaluated over the test signal. Therefore, the limen-based inequal- ity Eq.(2) is in no way related to the actual Gabor limit Eq.(1), and there is no expectation that the limen-based inequality should be satisfied. Secondly, one can straightforwardly use Fourier anal- ysis itself to "beat" Task 5, which again demonstrates that Task 5 does not test for violations of Fourier uncer- tainty since any Fourier-based analysis would necessarily be limited by the uncertainty principle. One method is to use a window Fourier Transform (WFT) to construct a spectrogram given by (cid:90) ∞ −∞ F (f, t0) = − (t−t0)2 ei2πf te 2γ2 X(t)dt (3) FIG. 1: F (f, t0) for the Gaussian pulse train using the pa- rameters Dt = 0.015s, Df = 4Hz, and σ = 50 ms. where X(t) is the pulse sequence and γ is the width of the window function. The γ is a free parameter that controls the aspect ratio of the individual signals as they appear in F (f, t0), and must be chosen to ensure that the signals do not overlap in F (f, t0). This can be readily achieved by setting γ equal to, for example, the temporal variance of the first pulse received in the pulse sequence. In Fig.1, we show F (f, t0) for the sequence of Gaussian pulses used in Ref.[2]. Fig. 1 clearly demonstrates that the function F (f, t0) can be used to obtain all the frequency and time shifts required to perform Task 5. When evaluating the WFT integral using a fast Fourier transform (FFT), the peak positions can be resolved to (at least) within one grid point in both frequency (df ) and time (dt). For a sampling rate R and time range T , the grid spacings in the FFT are given by dt = 1/R and df = 1/T . Therefore, the WFT would result in a limen-based uncertainty of δtδf = dtdf = 1/(RT ). Using R=96 kHz and T =2s, for example, we get δtδf = dtdf = 5.2083 × 10−6 which is orders of magnitude smaller than 1/(4π) ≈ 0.0796, and also orders of magnitude smaller than what was achieved by the human participants. Furthermore, since the WFT is a linear transform that can "beat" Task Time (s)Frequency (Hz)00.511.52350400450500550600650700750800 5, we also conclude that Task 5 does not test for the necessity of nonlinear transforms in models of human hearing. [2] J.N. Oppenheim and M.O. Magnasco, Phys. Rev. Lett., 110, 044301 (2013). [3] L. Cohen, Time-Frequency Analysis (Prentice Hall PTR, Englewood Cliffs, N.J., 1995). 2 [1] D. Gabor, Nature 159, 591 (1947).
1703.01357
7
1703
2019-09-27T21:22:14
A Computational Model of Systems Memory Consolidation and Reconsolidation
[ "q-bio.NC" ]
In the mammalian brain, newly acquired memories depend on the hippocampus for maintenance and recall, but over time the neocortex takes over these functions, rendering memories hippocampus-independent. The process responsible for this transformation is called systems memory consolidation. However, reactivation of a well-consolidated memory can trigger a temporary return to a hippocampus-dependent state, a phenomenon known as systems memory reconsolidation. The neural mechanisms underlying systems memory consolidation and reconsolidation are not well understood. Here, we propose a neural model based on well-documented mechanisms of synaptic plasticity and stability and describe a computational implementation that demonstrates the model's ability to account for a range of findings from the systems consolidation and reconsolidation literature. We derive several predictions from the computational model and suggest experiments that may test its validity.
q-bio.NC
q-bio
Running head: MODELING SYSTEMS RECONSOLIDATION Draft version 3.23, 2019-09-27. This paper has not been peer reviewed. Please do not copy or cite without author's permission. A Computational Model of Systems Memory Consolidation and Reconsolidation Peter Helfer1 and Thomas R. Shultz1 1McGill University, 2001 McGill College, Montreal, QC, Canada H3A 1G1 1 MODELING SYSTEMS RECONSOLIDATION Acknowledgements This work was supported by a Discovery Grant to TRS from the Natural Sciences and Engineering Research Council of Canada (7927-2012-RGPIN). The funding source had no other role than financial support. Abstract In the mammalian brain, newly acquired memories depend on the hippocampus for maintenance and recall, but over time the neocortex takes over these functions, rendering memories hippocampus-independent. The process responsible for this transformation is called systems memory consolidation. However, reactivation of a well-consolidated memory can trigger a temporary return to a hippocampus-dependent state, a phenomenon known as systems memory reconsolidation. The neural mechanisms underlying systems memory consolidation and reconsolidation are not well understood. Here, we propose a neural model based on well- documented mechanisms of synaptic plasticity and stability and describe a computational implementation that demonstrates the model's ability to account for a range of findings from the systems consolidation and reconsolidation literature. We derive several predictions from the computational model, and suggest experiments that may test its validity. Keywords: memory reconsolidation; artificial neural network; AMPA receptor exchange; neural plasticity; computational model 2 MODELING SYSTEMS RECONSOLIDATION A Computational Model of Systems Memory Consolidation and Reconsolidation The neural processes that transform memories from short-term to long-term storage are collectively known as memory consolidation. They include synaptic consolidation, relatively rapid intra-cellular changes that stabilize synaptic potentiation, and systems consolidation, slower and larger-scale processes that reorganize and restructure memory traces across brain systems. Specifically, systems consolidation refers to mechanisms that gradually make memories independent of the hippocampus, a structure in the medial temporal lobe of the mammalian brain. Whereas new memories are susceptible to disruption by a number of different types of interventions (e.g. electroconvulsive shock, certain pharmaceuticals, surgical procedures, and interference from new learning), consolidated memories are resistant to these treatments. However, retrieval of a consolidated memory can trigger a process in which it transiently becomes vulnerable to such interventions again, but subsequently restabilizes into a consolidated state. This is known as reconsolidation, and like consolidation it can be observed both at the synaptic and systems level. While much has been learned about the molecular underpinnings of synaptic consolidation and reconsolidation, the mechanisms responsible for the systems-level phenomena remain largely unknown (Asok, Leroy, Rayman, & Kandel, 2019; Hardt & Nadel, 2018). Here, we present an artificial neural network model that includes connection dynamics based on mechanisms of synaptic plasticity and demonstrate how these low-level processes can account for systems consolidation and reconsolidation. 3 MODELING SYSTEMS RECONSOLIDATION We begin with overviews of synaptic and systems memory consolidation and reconsolidation, and of previously published computer simulations. Next, we describe our model, report on simulation results and discuss their implications. Synaptic Consolidation and Reconsolidation Synaptic transmission. Neurons generate electrical signals called action potentials (APs) that travel along nerve fibers (axons) toward synapses where connections are made with other neurons. When an action potential reaches a synapse, neurotransmitter is released into the synaptic cleft, a narrow gap between the presynaptic active zone and the post-synaptic density (PSD), specialized areas of neuronal cell membrane that together make up the synapse. The neurotransmitter molecules bind to receptor proteins in the PSD, thereby triggering a response in the postsynaptic neuron. Depending on the type of neurotransmitter and the type of receptor, the response may be excitatory or inhibitory (making the postsynaptic neuron more or less likely to generate an AP), or have some other, e.g. regulatory, function. The average size of the excitatory or inhibitory response that is generated by the arrival of an action potential at a particular synapse is a measure of synaptic strength, and it depends both on the amount of transmitter released and on the numbers and types of receptors in the PSD (Citri & Malenka, 2008; Kandel, Dudai, & Mayford, 2014). Most neuroscientists believe that memories are stored in the strengths of synapses (Kandel et al., 2014; Sossin, 2008), an idea first articulated by Santiago Ramón y Cajal in the late 19th century (1894) and known as the synaptic theory of memory (Langille & Brown, 2018). Glutamate receptors. The amino acid glutamate is the most abundant neurotransmitter in the vertebrate nervous system (Platt, 2007). There are several types of glutamate receptors, among which the α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid receptor (AMPA 4 MODELING SYSTEMS RECONSOLIDATION receptor or AMPAR) is chiefly responsible for mediating excitatory synaptic transmission (Citri & Malenka, 2008; Maren, Tocco, Standley, Baudry, & Thompson, 1993), and the N-methyl-D- aspartate receptor (NMDA receptor or NMDAR) is involved with regulatory functions including the regulation of synaptic strength (Ben Mamou, Gamache, & Nader, 2006; Citri & Malenka, 2008; D. L. Walker & Davis, 2002). Long-term potentiation. When a neuron is stimulated strongly enough to make it fire (generate an AP), participating glutamatergic synapses are strengthened by a process called long- term potentiation (LTP) (Bliss & Lømo, 1973), which is associated with an increase in the number of AMPARs inserted in the PSD (Malenka & Bear, 2004). There are different stages of LTP. Moderately strong stimulation gives rise to early-phase LTP (E-LTP), which lasts for at most a few hours. More intense stimulation can trigger the induction of late-phase LTP (L-LTP), which can persist for months or longer (Abraham, 2003). Many researchers believe that LTP as studied in the laboratory is a cellular model for learning and memory, i.e. that it replicates the synaptic changes that occur during memory formation (Abraham, 2003; Bliss & Collingridge, 1993; Kandel, 2000; Lisman, 2017; Pastalkova et al., 2006), although conclusive proof for this has proved difficult to obtain (Lynch, 2004; Stevens, 1998). Consolidation. Induction of L-LTP -- but not E-LTP -- is believed to require RNA translation (synthesis of new proteins), based on experiments with protein synthesis inhibiting drugs (PSIs) such as anisomycin or cycloheximide. Infusion of such drugs before or immediately after stimulation can prevent establishment of L-LTP (Fonseca, Nägerl, & Bonhoeffer, 2006; Frey, Huang, & Kandel, 1993; Frey, Krug, Reymann, & Matthies, 1988; Huang, Li, & Kandel, 1994). However, once L-LTP has been established, a process that takes on the order of one hour, 5 MODELING SYSTEMS RECONSOLIDATION it is no longer vulnerable to PSI infusion (Bourtchouladze et al., 1998; Fonseca et al., 2006; Kandel, 2000). At the behavioral level, PSI injection within the first hour after training, but not later, has been shown to cause memory impairments (Davis & Squire, 1984; Judge & Quartermain, 1982; R. Kim, Moki, & Kida, 2011; Lattal & Abel, 2004), suggesting that the formation of long-term memory similarly requires protein synthesis (but see Canal and Gold (Canal, Chang, & Gold, 2007), who argue that, at least in the case of anisomycin, the memory impairments may be caused by other effects of the drug, not by protein synthesis inhibition). Reconsolidation. Several studies from the 1940s and 1950s demonstrated that electroconvulsive shock (ECS) could interfere with the establishment of long-term memory in rodents (Duncan, 1949; Gerard, 1955; Thompson & Dean, 1955) and in humans (Cronholm & Lagergren, 1959; Kehlet & Lunn, 1951), but only when applied within an hour or two after acquisition. However, in 1968, Misanin et al. reported that ECS could impair 24-hour-old, i.e. consolidated, memories of fear conditioning in rats - but only if the convulsive treatment was immediately preceded by memory "reactivation", i.e. retrieval cued by presentation of the conditioned stimulus (Misanin, Miller, & Lewis, 1968). Post-reactivation susceptibility to ECS was also demonstrated by Schneider and Sherman (1968) and Lewis, Mahan and Bregman (1972). Judge and Quartermain (1982) reported that injection of the protein synthesis inhibitor anisomycin, which was known to produce memory deficits when administered to mice immediately after training, could also impair older memories if given 30 minutes or less after reactivation. Przybyslawski and Sara (1997) showed that the NMDA receptor antagonist MK- 801 could induce memory deficits in rats that had been trained on a maze-running task, if injected up to 90 minutes after a maze run, but not after 120 minutes. The authors proposed that 6 MODELING SYSTEMS RECONSOLIDATION reactivation returns a well-established memory to a labile state from which it normally restabilizes spontaneously, and that this restabilization requires some or all of the same NMDA receptor-dependent events that are needed for consolidation of new memories. They therefore referred to the process as memory reconsolidation (Przybyslawski & Sara, 1997), a term first introduced by Spear (1973). Nader et al. (2000) demonstrated that anisomycin infusion into the amygdala of rats could disrupt an established fear-conditioning memory if performed immediately after reactivation, but not six hours later. Taken together, these studies support the notion that reactivation can render a memory trace that has undergone synaptic consolidation labile, and that an NMDA-dependent process, likely involving protein synthesis, is required to subsequently restabilize it. The phenomenon, known as synaptic memory reconsolidation, has attracted much interest in the wake of the Przybyslawski and Sara (1997) and Nader et al. (2000) papers, and a large reconsolidation literature now exists (for reviews, see Baldi & Bucherelli, 2015; Besnard, Caboche, & Laroche, 2012; Lee, Nader, & Schiller, 2017; Nader & Einarsson, 2010). A reconsolidation-like phenomenon has also been observed at the synaptic level: protein synthesis inhibition normally does not disrupt established (i.e. several hours old) L-LTP, but if administered together with low-frequency electric stimulation, it can cause depotentiation (Fonseca et al., 2006; Okubo-Suzuki et al., 2016). The function of reconsolidation. It is generally believed that the function of post- retrieval plasticity is to permit memory modification or updating when new information is encountered (Lee, 2009; Lee et al., 2017; Schiller & Phelps, 2011). Several studies with human subjects have shown that new training material is more likely to interfere with an established memory if presented after reactivation of the original memory (Forcato et al., 2007; Hupbach, 7 MODELING SYSTEMS RECONSOLIDATION Gomez, Hardt, & Nadel, 2007; M. P. Walker, Brakefield, Hobson, & Stickgold, 2003), although some authors have questioned whether such results really are evidence of a reconsolidation process (Ecker & Lewandowsky, 2012; Sederberg, Gershman, Polyn, & Norman, 2011). Post- reactivation memory updating has also been demonstrated in rodents (Jones, Pest, Vargas, Glisky, & Fellous, 2015; Monfils, Cowansage, Klann, & LeDoux, 2009). The molecular underpinnings of LTP. The biochemical basis of LTP is not completely understood, but intense research efforts during the last several decades have begun to throw light on some of the underlying molecular mechanisms (Baltaci, Mogulkoc, & Baltaci, 2019; Kandel et al., 2014). One significant discovery is that different types of AMPA receptors are of importance for the induction and maintenance of the early and late phases of LTP (Hong et al., 2013; McCormack, Stornetta, & Zhu, 2006; Rao-Ruiz et al., 2011). An AMPA receptor is made up of four subunits, each of which can be of several different kinds. Depending on its subunit composition, an AMPA receptor may or may not permit calcium ions to pass through the cell membrane, and is accordingly designated as calcium-permeable (CP) or calcium-impermeable (CI) (Henley & Wilkinson, 2013). E-LTP induction is characterized by a rapid increase in the number of CP-AMPARs, while the establishment of L-LTP requires insertion of CI-AMPARs (Clem & Huganir, 2010; Hong et al., 2013; Kessels & Malinow, 2009; Plant et al., 2006). This discovery is significant because the increased CP-AMPAR count is relatively short-lived, whereas an elevated level of CI-AMPARs can be sustained for a long time (Clem & Huganir, 2010; Hong et al., 2013; Plant et al., 2006). The ability of CI-AMPARs to remain at the PSD for a long time has been hypothesized to be due to molecular mechanisms that traffic CI-AMPARs to the potentiated synapse and protect them against removal (Clopath, Ziegler, Vasilaki, Buesing, & Gerstner, 2008; Helfer & Shultz, 2018; Jalil, Sacktor, & Shouval, 2015; Sacktor, 2011; 8 MODELING SYSTEMS RECONSOLIDATION Smolen, Baxter, & Byrne, 2012; Zhang, Smolen, Baxter, & Byrne, 2010). Interestingly, memory retrieval has been shown to trigger a transient reversal to a state with high CP-AMPAR and low CI-AMPAR counts (Hong et al., 2013; Rao-Ruiz et al., 2011), providing a potential explanation for post-retrieval synaptic instability. As discussed above, infusion of a protein synthesis inhibitor can interfere both with the induction of L-LTP and with restabilization of L-LTP after reactivation. A proposed explanation for this is that proteins required to transport CI-AMPARs into the synapse, and to maintain them there, need to be synthesized in order for either of these processes to occur (Sacktor, 2011). Additional evidence supporting the notion of LTP as a memory model includes other pharmacological interventions that affect both LTP and memory. For example, inhibition of NMDA receptors blocks both LTP and fear memory acquisition (Fanselow & Kim, 1994; Miserendino, Sananes, Melia, & Davis, 1990), inhibition of the protein kinase PKMζ with ZIP (zeta-inhibitory peptide) disrupts both L-LTP (Ling et al., 2002; Serrano, Yao, & Sacktor, 2005) and long-term memory (LTM) (Pastalkova et al., 2006; Shema, Hazvi, Sacktor, & Dudai, 2009; Shema, Sacktor, & Dudai, 2007), and blocking endocytosis (removal from the cell membrane) of CI-AMPARs rescues both L-LTP and LTM from these effects of ZIP (Migues et al., 2010, 2016). Also, optogenetic stimulation protocols that induce long-term depression (LTD) and LTP have been shown to inactivate and reactivate, respectively, a conditioned fear response (Nabavi et al., 2014). To summarize, LTP is considered a cellular model of memory, because it is compatible with the synaptic theory of memory, and although a causal relationship has proved difficult to establish, many results support the notion that the synaptic changes that characterize LTP also play an important role in the cellular mechanisms of memory. 9 MODELING SYSTEMS RECONSOLIDATION Systems Consolidation and Reconsolidation Consolidation. Scoville and Milner (1957) famously documented that bilateral hippocampal lesions in human patients resulted in profound memory loss for past events (retrograde amnesia), as well as a nearly complete inability to form new long-term memories (dense anterograde amnesia). The retrograde amnesia appeared to be graded: it was most severe for the period shortly before surgery, while older memories were relatively spared. These findings have been confirmed with other human patients (Dede & Smith, 2016; Manns, Hopkins, & Squire, 2003; Penfield & Milner, 1958), and reproduced in animal studies with primates (Squire & Zola-Morgan, 1991; Zola-Morgan & Squire, 1990) and rodents (Anagnostaras, Maren, & Fanselow, 1999; Winocur, 1990) (although a temporal gradient is not always observed (Lah & Miller, 2008; Sutherland, Sparks, & Lehmann, 2010)). The observation that recent memories are more vulnerable to hippocampal damage than older ones has given rise to the notion that the hippocampus plays a crucial role in the establishment, maintenance and recall of new memories, but over time memories become hippocampus-independent (Frankland & Bontempi, 2005; McClelland, McNaughton, & O'Reilly, 1995; Milner, 1989; Nadel & Hardt, 2010; Squire & Zola-Morgan, 1991). The process responsible for this putative reorganization is called systems memory consolidation (Dudai, 2004). According to what is called the standard theory of systems consolidation (Squire & Alvarez, 1995), the hippocampus quickly records information about events in real time, then repeatedly replays them, perhaps primarily during sleep (Genzel et al., 2017), thereby driving a more time-consuming process of memory trace creation in the neocortex (Frankland & Bontempi, 2005; McClelland et al., 1995; Teyler & Rudy, 2007). A possible explanation for why it takes longer to lay down a memory trace in neocortex is that it is more sparsely inter-connected than the hippocampus, and that creation of a memory trace therefore 10 MODELING SYSTEMS RECONSOLIDATION requires axonal growth and synaptogenesis (Alvarez & Squire, 1994; Frankland & Bontempi, 2005). Another possibility, suggested by Teyler and Rudy (2007), is that strong inhibitory control mechanisms in the neocortex slow down the induction of LTP. The idea that hippocampal "replay" can trigger reinstatement of activation patterns in neocortex raises the question of what kind of information is stored in the hippocampus. McClelland et al., in their "two memory systems" proposal (1995), suggested that the hippocampus stores "compressed representations" of cortical patterns, that can be decoded to reconstruct activation patterns in neocortex, but most authors favor a view where the hippocampus does not store memory content, but rather "indices" (Teyler & Discenna, 1986; Teyler & Rudy, 2007) or "links" to loci in the neocortex where components of memory traces reside (Alvarez & Squire, 1994; Frankland & Bontempi, 2005). In these accounts, the hippocampus plays a dual role: it supports, through replay, the establishment and/or strengthening of intra-neocortical connections, and it also assist in memory retrieval until those connections become strong enough to function independently. Not all researchers agree with the standard theory of systems consolidation. In particular, proponents of the Multiple-Trace Theory (Nadel & Moscovitch, 1997) and its descendant the Trace Transformation Hypothesis (Moscovitch, Cabeza, Winocur, & Nadel, 2016; Nadel, Winocur, Ryan, & Moscovitch, 2007; Winocur, Sekeres, Binns, & Moscovitch, 2013) hold that whereas semantic memories (knowledge about the world) are consolidated in the manner described by the standard theory, autobiographical memories (spatial and temporal information) remain hippocampus-dependent for as long as they persist. According to this view, the loss of hippocampal support (whether due to lesion/damage or to gradual weakening as memories age) results in the loss of spatio-temporal information. Empirical support for the notion that spatial 11 MODELING SYSTEMS RECONSOLIDATION specificity is lost as memories age includes rodent studies by Wiltgen et al. (2010), Gafford, Parsons, & Helmstetter (2013) and Einarsson et al. (2015). On the other hand, Dede and Smith, in a comprehensive review of human cases of temporal lobe damage (Dede & Smith, 2016), conclude that there is no evidence for loss of remote episodic information after hippocampal damage, contradicting the transformation hypothesis and supporting the standard theory. While the standard theory thus is not universally accepted (Nadel & Moscovitch, 1997; Rudy, Biedenkapp, & O'Reilly, 2005; Winocur, Moscovitch, & Bontempi, 2010), it enjoys widespread support and is compatible with a large body of empirical evidence (Dede & Smith, 2016; Runyan, Moore, & Dash, 2019; Squire, Genzel, Wixted, & Morris, 2015; Wang & Morris, 2010; Wiltgen, Brown, Talton, & Silva, 2004). Several neocortical regions in the frontal and temporal lobes have been identified as locations where memories consolidate, including the frontal cortex, posterior and anterior cingulate and temporal cortex (Bontempi, Laurent-Demir, Destrade, & Jaffard, 1999; Frankland, Bontempi, Talton, Kaczmarek, & Silva, 2004; Quirk & Mueller, 2008), parietal and retrosplenial cortex (Maviel, Durkin, Menzaghi, & Bontempi, 2004) (both using a spatial discrimination task in rodents), and frontal, lateral temporal, and occipital lobes for autobiographical memories in humans (Bayley, Gold, Hopkins, & Squire, 2005). For reviews, see Dede and Smith (2016) and Frankland and Bontempi (2005). The anterior cingulate cortex (ACC), part of the prefrontal cortex, has received particular attention because it has been shown to play an important role for recall of remote memories in several experimental paradigms including fear conditioning, conditioned taste aversion, trace eyeblink conditioning and spatial discrimination (Bontempi et al., 1999; Ding, Teixeira, & Frankland, 2008; Einarsson & Nader, 2012; Einarsson et al., 2015; 12 MODELING SYSTEMS RECONSOLIDATION Frankland & Bontempi, 2005; Frankland et al., 2004; Maviel et al., 2004; Takehara, Kawahara, & Kirino, 2003). Experimental methods. Much information regarding the role of the hippocampus in systems consolidation comes from lesion experiments with rats or mice, where bilateral hippocampal lesions are performed at different intervals following training. In a common type of fear conditioning, a sound or a specific spatial context (conditioned stimulus) is paired with an electric foot shock or other aversive unconditioned stimulus. Recall is subsequently tested by presenting the conditioned stimulus alone and measuring the degree of fear response. Findings from this type of study indicate that hippocampal lesions immediately after training result in severely impaired recall, but longer delays between conditioning and lesions produce gradually less severe impairments. The time interval after conditioning during which hippocampal lesions result in significant memory impairment is called the systems consolidation window. For fear conditioning in rodents, most studies report a consolidation window of between three and four weeks (J. Kim, Clark, & Thompson, 1995; J. Kim & Fanselow, 1992; Kitamura et al., 2009; Land, Bunsey, & Riccio, 2000; Winocur, Moscovitch, & Sekeres, 2013). Longer windows have been reported for primates: months in monkeys (Zola-Morgan & Squire, 1990), and years in humans (Scoville & Milner, 1957; Squire & Cohen, 1979). Systems consolidation has also been investigated by studying the effect of reversible inactivation of specific brain areas. Studies using pharmaceutical inactivation of the rodent hippocampus and/or ACC have shown that retrieval of a fear memory is hippocampus-dependent one or three days after acquisition, but not after 28 or 30 days. At this point it has instead become dependent on the ACC for retrieval (Einarsson et al., 2015; Frankland et al., 2004; Sierra et al., 2017; Wiltgen et al., 2010). 13 MODELING SYSTEMS RECONSOLIDATION Thus results from both lesion and inactivation studies are compatible with the standard theory, and similar results have also been obtained in studies measuring brain activity during retrieval of recent and remote memories (Bontempi et al., 1999; Gafford et al., 2013; Nadel & Hardt, 2010). Reconsolidation. Whereas synaptic reconsolidation -- the transient post-reactivation susceptibility of memories to amnestic interventions like electro-convulsive shock or administration of protein synthesis inhibitors -- has been studied since the 1960s (Judge & Quartermain, 1982; Lewis et al., 1972; Misanin et al., 1968), systems reconsolidation -- the return of hippocampus-dependence after reactivation -- was first described by Land et al. in 2000 (Land et al., 2000), and subsequently by Debiec et al. (2002) and Winocur et al. (2009; 2013). These researchers found that hippocampal lesions produced amnesia for 30- or 45-day old fear memories if performed immediately after reactivating the memories by presentation of the conditioned stimulus. Hippocampal lesions did not produce memory impairments without preceding reactivation, nor if administered after the reactivated memory was allowed 48 hours to restabilize after reactivation (Debiec et al., 2002). It thus appears that reactivation renders a neocortical memory trace unstable and that a functioning hippocampus is needed for its restabilization. Inactivation studies have provided additional information about the effect that reactivation has on remote memories. As noted above, 30-day old fear memories are strongly dependent on ACC for retrieval. In an elegant study, Einarsson et al. (2015) showed that six hours after memory reactivation, ACC inactivation no longer impaired retrieval, but after 24 hours ACC-dependence had returned. At the six-hour time point, inactivation of the 14 MODELING SYSTEMS RECONSOLIDATION hippocampus also did not affect retrieval, but simultaneous activation of both hippocampus and ACC did block recall. Together, the findings from post-reactivation lesion and inactivation studies suggest that reactivation triggers the creation of a short-lived hippocampal memory trace that is able to support recall and is also required for restabilization of the ACC trace. An additional interesting finding is that PSI infusion into hippocampus immediately after reactivation blocks reconsolidation when performed three days (Debiec et al., 2002; 2006), five days (Rossato, Bevilaqua, Medina, Izquierdo, & Cammarota, 2006), or seven days (Haubrich et al., 2015; Milekic & Alberini, 2002) after training. These results suggest that recovery from the destabilization caused by reactivation depends on protein synthesis in the hippocampus. Researchers differ regarding the maximum memory age at which reactivation can trigger systems reconsolidation. Whereas the lesion and inactivation studies indicate that memories are susceptible for at least a month after training (Debiec et al., 2002; Einarsson et al., 2015; Land et al., 2000; Winocur et al., 2009; Winocur, Sekeres, et al., 2013), the results from post-reactivation PSI infusion into the hippocampus are less clear. Debiec et al. (Debiec et al., 2002) were able to demonstrate reconsolidation blockade with anisomycin as late as 45 days after training, whereas Frankland et al. (Frankland et al., 2006) were unable to show the effect at 36 days. The reason for this discrepancy is not clear. An interesting clue to a connection between synaptic and systems reconsolidation is provided by Ghazal (2016): reactivation of a 1-day-old fear memory triggered a strong reduction of CI-AMPARs in hippocampus but not in ACC, whereas at 30 days the opposite was true. This result is consistent with a transition from engagement of the hippocampus to the ACC over a 15 MODELING SYSTEMS RECONSOLIDATION systems-consolidation timeframe and with synaptic reconsolidation taking place in whichever of the two systems is engaged in the retrieval. In summary, retrieval of a new (e.g. 3-day-old) fear memory requires the hippocampus but not the ACC. Over time, a reversal takes place so that retrieval of a 30-day-old memory requires the ACC but not the hippocampus. Reactivation of a consolidated memory temporarily returns it to ACC-independence for retrieval. Systems consolidation (establishment of a neocortical trace) and systems reconsolidation (restabilization after reactivation-induced destabilization of the neocortical trace) both require hippocampal involvement. Model Based on the findings described in the foregoing, we here present a model of systems memory consolidation and reconsolidation. Although the role of LTP in learning and memory remains conjectural, for the purpose of modeling we assume that LTP -- or an LTP-like mechanism -- is the cellular substrate of memory and investigate to what extent this type of synaptic mechanism can explain the systems-level phenomena. Below we provide a conceptual description of the model; the computational implementation is described in the Methods section. Synaptic level • Moderately intense stimulation induces E-LTP, which involves the rapid insertion of CP-AMPARs. Constitutive processes subsequently remove these within hours. • More intense stimulation sets in motion L-LTP induction (synaptic consolidation) which involves a state change in a bistable mechanism (molecular switch). When in the ON state, this mechanism maintains a high CI-AMPAR count in the synapse; 16 MODELING SYSTEMS RECONSOLIDATION when the switch is OFF, the CI-AMPAR count drifts towards the basal level characteristic of the unpotentiated synapse. • Memory retrieval abruptly removes CI-AMPARs from the synapse and replaces them with CP-AMPARs, thus returning the synapse to an E-LTP-like state. The subsequent restoration of L-LTP is protein-synthesis-dependent and requires NMDAR activity, i.e. neural stimulation. Systems level As discussed above, some researchers disagree with the standard theory, in particular with respect to episodic/autobiographical memories. Nevertheless, there is widespread agreement that semantic information does undergo systems consolidation. The aim of the present work is to model those types of memories that do undergo a reorganization whereby they gradually become hippocampus-independent; the question of whether episodic/autobiographical memories are processed differently is beyond our present scope. • Stimulus presentations trigger patterns of activation in multiple ensembles of neurons in the neocortex (NC). These active neurons in turn project onto and activate neurons in the hippocampus (HPC), where a memory trace is quickly created, providing linkages between the activated NC ensembles. Linkages are also created through prefrontal cortex, in particular the ACC, but these connections are initially too weak to support memory retrieval without HPC support. • Subsequently, the HPC memory trace is spontaneously and repeatedly reactivated which causes stimulation of these same NC neural ensembles through nerve fibers projecting back from the HPC to the NC. Over time, the repeated reactivation of the 17 MODELING SYSTEMS RECONSOLIDATION NC neural ensembles strengthens the ACC linkages, eventually to a point where they can support retrieval of the memory without assistance from the hippocampus. • Meanwhile, the HPC trace is gradually weakened by constitutive decay processes (Hardt, Nader, & Nadel, 2013; Sachser et al., 2016). • If a consolidated memory is reactivated by a reminder, then activity in the neocortical neural ensembles triggers re-establishment of the HPC linkage. Simultaneously, the retrieval causes transient destabilization of the synapses in the ACC linkage (synaptic reconsolidation). • Following reactivation, hippocampal replay stimulates the now destabilized synapses of the ACC linkage. This activity drives restabilization of these synapses. Meanwhile, the reactivated HPC trace rapidly decays, leading to a return to ACC- dependence in 24 hours or less. Temporal characteristics The temporal characteristics of systems consolidation and reconsolidation provide some clues to the underlying mechanisms: The reconsolidation window is shorter than the consolidation window. Hippocampal lesion after training produces a memory impairment, more severe the shorter the delay between conditioning and lesion. Similarly, hippocampal lesion after reactivation also causes more severe impairment if performed early. However, the time scales are very different. New memories are vulnerable for at least several weeks after training (J. Kim & Fanselow, 1992; Squire & Cohen, 1979; Winocur, 1990; Zola-Morgan & Squire, 1990), whereas the post-reactivation window of vulnerability only lasts for a few days (Debiec et al., 2002). Our model attributes this 18 MODELING SYSTEMS RECONSOLIDATION difference to the different natures of the processes being interrupted by HPC lesioning in the two scenarios. In the case of consolidation, what is being interrupted is the gradual and relatively slow establishment and strengthening of intra- neocortical connections; thus the earlier the intervention is performed, the weaker the partially consolidated memory trace will be. Reconsolidation blockade, on the other hand, interrupts the much faster process of re-stabilization of destabilized neocortical synapses. The earlier in the reconsolidation window hippocampal lesion is performed, the fewer synapses will have had time to restabilize, leading to more severe memory loss. Memories become transiently ACC-independent after reactivation. Whereas a consolidated fear memory depends strongly on the ACC for recall, reactivation triggers a brief period of ACC-independence, such that six hours after reactivation ACC inactivation has little or no effect on retrieval, but 24 hours after reactivation full ACC dependence has returned (Einarsson et al., 2015). This finding suggests that reactivation triggers the creation of a short-lived hippocampal memory trace that is able to support recall six hours later, but not 24 hours later. For a memory trace to last for six hours or more, it must have undergone synaptic consolidation, yet 24 hours is a short lifetime for a consolidated trace -- compare the situation after initial acquisition, where the hippocampal trace is able to support recall for at least several days. The mechanism that causes a faster decline of a post- reactivation hippocampal trace is unknown. 19 MODELING SYSTEMS RECONSOLIDATION Previously Published Computational Models Several computational models of systems consolidation have been published. Alvarez and Squire (1994) described a recurrent artificial neural network (ANN) model with three memory banks representing, respectively, the medial temporal lobe (MTL - the hippocampus and adjacent areas) and two areas of neocortex (NC). Learning was implemented by a Hebbian-like algorithm that adjusted connection weights, with a fast learning rate for the connections between the MTL and each of the NC areas (MTL-NC connections) and a slower rate for the connections between the two NC areas (NC-NC connections). After training on a pair of random patterns, one in each NC area, the model could be presented with either of the two patterns and reproduce the other with good accuracy thanks to the strong linkage provided by the MTL-NC connections, but not if the MTL subsystem was disabled. However, if after initial training, MTL units were repeatedly reactivated in random fashion, then the resulting stimulation through the strong MTL- NC connections would cause the two trained NC patterns to become active, and Hebbian learning would strengthen the NC-NC connections to a point where they eventually became able to support recall without the MTL linkage. The Alvarez and Squire model thus demonstrated the computational feasibility of the standard theory of systems consolidation. McClelland et al. (1995) hypothesized that systems consolidation serves to avoid "catastrophic interference", impairment of older memory traces when new patterns are integrated too quickly, an effect that had been observed in artificial neural networks. They suggested that the brain avoids this problem by having the hippocampus quickly capture new patterns, then repeatedly replaying them to the slow-learning neocortex. A computational model described in the paper demonstrated that such repeated replay, interleaved with rehearsal of previously 20 MODELING SYSTEMS RECONSOLIDATION learned patterns, could avoid catastrophic interference. The model used in their simulations consisted of a traditional three-layer feed-forward network. Meeter and Murre's TraceLink model (Meeter & Murre, 2005; Murre, 1996) combines elements from these two earlier studies. Like Alvarez and Squire's (1994) model, TraceLink is a recurrent network with neocortical and hippocampal layers. Unlike the Alvarez and Squire model, it is capable of learning multiple patterns in succession, avoiding catastrophic interference by repeated activation of previously learned material. It differs from the McClelland et al. (1995) model by including simulation of a hippocampal system which autonomously presents a random mix of new and old patterns to neocortex. McClelland et al.'s simulation did not include a hippocampus component; the interleaved set of old and new training patterns was provided to the network by the programmer. A paper by Amaral et al. (2008) is primarily concerned with demonstrating that a single model can accommodate seemingly contradictory results with respect to the transience or permanence of amnesia resulting from hippocampal lesions at different time points in the course of systems consolidation. The model consists of two recurrent layers, representing hippocampus and neocortex and employs the "synaptic reentry reinforcement" (SRR) mechanism, first described by Wittenberg and Tsien (2002), which can autonomously strengthen a memory trace over time, and even repair partial damage. The model is able to reproduce the empirical finding that both early and late hippocampal disruption can produce either permanent or transient memory impairment, but permanent loss is a more likely outcome after early and/or extensive hippocampal disruption. These simulations thus illuminate various aspects of systems consolidation. Our work builds on these previous models and adds the capability to simulate systems reconsolidation, i.e. 21 MODELING SYSTEMS RECONSOLIDATION a transient period of hippocampus-dependence following cued retrieval of a consolidated memory. The key to this capability is a more elaborate connection design than is traditionally used in neural networks, Specifically, the connections in our model simulate the AMPA receptor exchanges underlying synaptic consolidation and reconsolidation. Below, we introduce our computational implementation of the model described above, and show that it is capable of reproducing both systems consolidation and reconsolidation, including the effects of different pharmaceutical and surgical interventions. 22 MODELING SYSTEMS RECONSOLIDATION Methods Simulation Targets Table 1 summarizes the empirical findings, as described in the introduction, which our model aims to reproduce in simulations. Table 1: Simulation targets 3 PSI during conditioning impairs LTM but not STM.5,7,8 4 PSI infusion during maintenance does not cause memory impairment.5,7 5 HPC lesion at 3d causes memory impairment.9,10 HPC lesion at 3 days or less after training causes Result 1 Retrieval is HPC-dependent and ACC- independent at 3d.1,2,3,4,5 2 Retrieval is ACC-dependent and HPC- independent at 30d.1,2,3,4,6 6 HPC lesion at 30d does not cause memory impairment.9,10,11 7 Reactivation alone does not impair a consolidated memory.10,11 8 Reactivation + PSI causes memory impairment.8,11,12,13,14 9 Delayed impairment effect of post-reactivation PSI infusion in HPC.8,11 10 Reactivation + HPC lesion causes memory impairment.10,11,15 11 Graded effect of post-reactivation HPC lesion.11 12 Retrieval can be supported by either ACC or HPC 6h after reactivation.1 Description Retrieval of a three-day old memory is impaired by HPC inactivation but unaffected by ACC inactivation. Retrieval of a 30-day old memory is impaired by ACC inactivation but unaffected by HPC inactivation. Systemic injection of PSI during training prevents LTM induction -- but does not impair STM . Systemic PSI injection does not impair a consolidated memory. later recall to be severely impaired compared to non- lesioned animals. HPC lesion, when not preceded by memory reactivation, does not result in subsequently impaired recall. Reactivation without subsequent HPC lesion or PSI infusion does not by itself impair a consolidated memory. Reactivation immediately followed by PSI infusion into HPC causes impairment of a consolidated memory. The impairment caused by post-reactivation PSI infusion in HPC does not manifest in retrieval test soon after lesion (4h) but only later (24h). Reactivation immediately followed by HPC lesion causes impairment of a consolidated memory. The severity of memory impairment caused by post- reactivation HPC lesion diminishes with increasing reactivation-lesion delay. Six hours after reactivation, neither ACC inactivation nor HPC inactivation alone impairs retrieval, but simultaneous inactivation of both does block retrieval. 1Einarsson et al. (2015) 2Gafford et al. (2013) 3Ghazal (2016) 4de Oliveira Alvares et al. (2012) 5Frey (1997) 6Frankland et al. (2004) 7Davis and Squire (1984) 8Rossato et al. (2006) 9J. Kim and Fanselow (1992) 10Land et al. (2000) 11Debiec et al. (2002) 12Frankland et al. (2006) 13Haubrich et al. (2015) 14Milekic and Alberini (2002) 15Winocur et al. (2009) 23 MODELING SYSTEMS RECONSOLIDATION Model Design Like other artificial neural networks, ours consists of units and connections, where units are analogs of biological neurons (or ensembles of neurons) and connections model synapses. (For a primer on artificial neural networks, see e.g. Shultz (2003)). Topology. We use a recurrent artificial neural network with four regions representing hippocampus (HPC), anterior cingulate cortex (ACC) and two sensory cortex areas, SC0 and SC1, to which stimuli are presented. Each region consists of 25 units. The units are linked together by connections, such that each HPC and ACC unit is bidirectionally connected to all the units in the other three regions. The set of connections between any two regions, e.g. from HPC to SC1, is referred to as a tract, see Figure 1. 24 MODELING SYSTEMS RECONSOLIDATION Figure 1: Network architecture. To reduce clutter only three of the twenty-five units in each region are shown. Each double-headed arrow represents two independent connections, one in each direction, between a pair of units. The diagram illustrates the state after initial acquisition: presentation of the unconditioned stimulus (US) and the conditioned stimulus (CS) has activated some units in SC0 and SC1 (filled circles) and fast learning has created strong linkages (bold lines) through the HPC. Linkages through the slower-learning ACC are still weak. This topology is an evolution of the one used by Alvarez and Squire in their influential model of systems memory consolidation (Alvarez & Squire, 1994). Their model had two neocortical regions with direct linkage between them; we have added the ACC region in order to be able to simulate interruption of linkage between the SC1 and SC2 regions by reversible inactivation of either the ACC and/or HPC regions (or both). As noted in the introduction, a number of neocortical regions other than the ACC have been identified as loci for remote memory traces. For simplicity, the model only includes the ACC as a target region for systems consolidation. Similarly, the model's connectivity between the hippocampus and neocortical regions is simplified for reasons of computational feasibility. In a real brain, the hippocampal complex is comprised of many intricately interconnected subareas that receive input from and project back to not only sensory cortices, but to a widespread network of different structures in hierarchies of association areas. For an overview, see Lavenex and Amaral (2000). The algorithms for activation, learning, and retrieval are adapted from Meeter and Murre's TraceLink model of systems memory consolidation (Meeter & Murre, 2005), with minor modifications that will be detailed in the following. 25 MODELING SYSTEMS RECONSOLIDATION The model description that follows makes reference to a number of configurable parameters. These are documented in tables 2-4. Units. The units are bistable and stochastic. A unit's activity level is either 1 (active) or 0 (inactive). The probability that a unit is active (the activation function) is an asymmetric sigmoid function of its net input: ( P isActive ) j = 1 ( actK net − inhib r j ) 1 + e− Where j P isActive is the probability that unit j is active, and j ) ( (1) jnet is the net input to unit j, calculated as the sum of the activity levels of units connected to unit j, weighted by inbound connection strengths: net j = ∑ i a s i ij Here, ia is the activation level of unit i and ijs is the strength of the connection from unit i to (2) unit j. The actK symbol in Equation 1 is a configurable parameter that controls the slope of the sigmoid function. The inhib variable in Equation 1 is a region-specific dynamic inhibition level, r which will be described below. Figure 2 shows two examples of activation functions with actK values of 2.0 and 10.0, respectively, both with an inhibition level of 5.0. The probability of activation grows from near zero when net input is low to near one when net input is high, with a value of 0.5 at the current inhibition level. An increase in inhibition shifts the curve to the right, with the effect that greater net input is required for activation. This tends to results in fewer active units in the region. Greater actK values produce curves with steeper, more abrupt transitions, corresponding to a 26 MODELING SYSTEMS RECONSOLIDATION more deterministic behavior. All simulations reported here use actK = 2.0. Figure 2: Activation functions. Two examples are shown, one with actK = 2.0 (solid) and one with actK = 10.0 (dashed). Both are plotted with inhibr = 5.0. The use of the activation function, including the inhibition mechanism, will be further described below. Connections. The connections are abstract models of glutamatergic synapses, characterized by four variables: • psdSize: the maximum number of AMPARs that can be inserted, analogous to the number of receptor "slots" in the PSD. • numCpAmpars: the number of currently inserted CP-AMPARs • numCiAmpars: the number of currently inserted CI-AMPARs • isPotentiated: a Boolean attribute that models the bistable nature of L-LTP This combination of binary and continuously variable attributes makes it possible to model synapses where potentiation behaves like a bistable switch (Bortolotto, Bashir, Davies, & Collingridge, 1994; Jalil et al., 2015; Westmark et al., 2010), yet synaptic strength is variable with many distinct levels (Bartol et al., 2015). 27 MODELING SYSTEMS RECONSOLIDATION As will be described below, stimulation causes an increase of the psdSize attribute, allowing more AMPARs to be inserted. A connection's weight at any moment is proportional to its total number of inserted AMPARs (numCpAmpars + numCiAmpars). Simulation A simulation consists of a sequence of time steps. Various interventions may be scheduled for any time point during the simulation, and in addition several background processes execute at each time step. The scheduled event types are training, reactivation, HPC lesion, HPC inactivation and ACC inactivation. The background processes are consolidation, AMPAR trafficking and random depotentiation. In addition, a retrieval test can be executed at any time. The different interventions and background processes will be described below, but first we introduce the concept of a learning cycle. Learning cycle. Learning takes place when stimuli are presented, when a memory is reactivated, and during memory consolidation. In each of these instances, learning is achieved by the execution of a learning cycle in every connection that satisfies the Hebbian condition that the two units that it connects are both active. A learning cycle consists of three steps: PSD growth, CP-AMPAR influx, and (possibly) L-LTP induction. 1. PSD growth. The network learns activation patterns by a Hebbian learning rule (Hebb, 1949) that increases the psdSize attribute of connections between simultaneously activated units, asymptotically towards a maximum value: psdSize ij ← psdSize ij + µ T ( maxPsdSize − psdSize ij ) (3) where psdSize is the PSD size of the connection between units i and j, maxPsdSize is a ij configurable parameter, and Tµ is a learning rate specific to the tract that connection ij belongs to. This learning rule differs from that of TraceLink in three respects: (1) We use an asymptotic 28 MODELING SYSTEMS RECONSOLIDATION function where TraceLink used a linear function with hard limits; this is an esthetic choice; (2) In TraceLink, the learning rule applies directly to connection weights; in our model, learning affects PSD size, i.e. the maximum weight that the connection can attain; actual weight is defined by the number of inserted AMPARs, which may at times be lower than psdSize; (3) TraceLink's learning rule supported "unlearning" (learning at a negative rate) to model interference, a phenomenon that we are not addressing in the present model. To simulate different intensities of stimulation, the learning rule is applied a variable number of times, specified by the parameter numStimCycles. The stimulation resulting from spontaneous hippocampal activation during consolidation is modeled as much less intense than the stimulation due to stimulus presentation. 2. CP-AMPAR influx. PSD growth is followed by an increase in the number of CP-AMPARs such that the total AMPAR count becomes equal to the connection's PSD size: numCpAmpars ij ← psdSize ij − numCiAmpars ij (4) This models the rapid CP-AMPAR influx during E-LTP induction. 3. L-LTP induction. Finally, probabilistic induction of L-LTP in a connection is implemented by turning on its isPotentiated attribute with a probability that depends on the intensity of the stimulation: prob ← f ( numStimCyc les , stimTh resh ) (5) where numStimCycles is as described above, stimThresh defines the value of numStimCycles that produces a 0.5 probability of L-LTP induction, and f is an asymmetric sigmoid function: ( f stim thresh , ) = 1 stim thresh − ) 1 + ( e− (6) 29 MODELING SYSTEMS RECONSOLIDATION Learning cycles are executed (a) when stimuli are presented for training, (b) at memory retrieval (reactivation), and (c) when patterns are spontaneously activated by the memory consolidation process. Descriptions of these mechanisms follow: Training. To train an association, patterns (subsets) of units are activated in the SC0 and SC1 regions to represent an unconditioned stimulus, US, and a conditioned stimulus, CS, respectively. Each stimulus pattern consists of k *r numUnits units, where r numUnits is the r number of units in region r and rk is a region-specific parameter that specifies the fraction of the region's units to activate in a pattern. (This is described further in the section about inhibition, below.) The network then randomly selects and activates rk linkage units in each of the HPC and ACC regions and executes a learning cycle in all connections that connect two simultaneously active units, i.e. connections that are in the Hebbian condition. The learning rate, Tµ , is defined to be relatively high in HPC tracts, allowing rapid creation of linkages strong enough to support recall. The ACC learning rate is lower, hence linkages through the ACC region are not strong enough to independently support recall immediately after training. Retrieval. To test recall of a trained pattern, the CS units are activated in SC0, and the network is cycled by repeated application of the activation function in all units a number of times specified by the parameter numSettleCycles. The units are updated synchronously: new activation values are first computed for all units, based on their current net inputs. The new values are then applied to all the units. The activity pattern that the SC1 region eventually settles into is compared to the associated US pattern to calculate a retrieval test score. Inhibition. To ensure fast settling of the network, we use a variation of an inhibition mechanism described by Meeter and Murre for their TraceLink architecture, designed to 30 MODELING SYSTEMS RECONSOLIDATION (roughly) model the working of inhibitory neurons (Meeter & Murre, 2005). Each region has a configurable target fraction of active units, rk . Following each update cycle during the settling phase of memory retrieval (and during consolidation, see below), each region's inhibition level, inhib (see Equation 1), is adjusted by an amount proportional to the current deviation of the r number of active units in the region from the target value defined by the region's rk value: inhib r ← inhib r + numActive r r k − k r * numUnits r * inhibIncr inhib r ← ( min maxInhib max minInhib inhib r ( ( , ))) (7) (8) where numActive is the number of active units in region r and r numUnits is the total number of r units in region r. The configurable inhibIncr parameter controls the rate of adjustment of the inhibition level. The assignment in Equation 8 restricts inhib to a configurable range of values r [ ] minInhib maxInhib . , Systems consolidation. At every simulation time step a randomly selected trained pattern is activated in HPC, after which the entire network is cycled in the same manner as for recall test (but without stimulus presentation). Whatever pattern the ACC and SC regions settle into is then reinforced by application of the learning rule. Because the network is more likely to settle into trained patterns than other random states, this will tend to strengthen CS-US linkages through the ACC, eventually making recall of trained patterns HPC-independent. AMPAR trafficking. At each simulation time step, AMPAR trafficking is simulated in all connections as follows: CP-AMPARs: the numCpAmpar value declines asymptotically towards zero, simulating the limited dwell time of CP-AMPARs at the synapse: 31 MODELING SYSTEMS RECONSOLIDATION numCpAmpars ij ← numCpAmpars ij − cpAmparRemovalRate *( numCpAmpars minNumCpAmpars − ij where numCpAmpars is the current number of CP-AMPARs at connection ij, and ij cpAmparRemovalRate and minNumCpAmpars are configurable parameters. (8) ) CI-AMPARs: if a connection's isPotentiated attribute is true and its source and destination units are both active (Hebbian condition), then numCiAmpars grows asymptotically towards the number of available slots in the connection (psdSize), simulating activity-dependent trafficking of CI-AMPARs into a potentiated synapse: numCiAmpars ij ← ( min numCiAmpars ij + ciAmparInsertionRate psdSize ij , ) (9) In unpotentiated connections, numCiAmpars declines exponentially, simulating a depotentiated ij synapse's gradual return to baseline: numCiAmpars ij ← numCiAmpars ij − ciAmparRemovalRate *( numCiAmpars minNumCiAmpars − ) (10) minNumCiAmpars, ciAmparInsertionRate and ciAmparRemovalRate are configurable parameters. Depotentiation. Potentiated connections are subject to random constitutive depotentiation at a rate controlled by the tract-specific parameter baseDepotProb ,which has a t higher value for HPC tracts than for intra-neocortical tracts, to account for the observed faster decline of hippocampal memory traces. Reactivation. Reactivation is modeled as an unreinforced CS presentation, i.e. a cued retrieval. The CS pattern is activated in the SC0 region, the network is settled as described for training, and an AMPAR exchange is simulated in all connections situated between simultaneously active units (i.e. connections in the Hebbian condition): numCiAmpars is ij reduced to its minimum value minNumCiAmpars, and numCpAmpars is set to its maximum ij 32 MODELING SYSTEMS RECONSOLIDATION allowed value numCiAmpars ij psdSize ij − . This puts those consolidated ACC linkage connections that were activated by memory retrieval in an unstable E-LTP-like state, modeling the post-reactivation instability documented in empirical studies. A random set of rk HPC linkage units is then activated, in the same manner as during initial acquisition, and a round of Hebbian learning takes place. As noted in the introduction, the hippocampal engagement after reactivation is much briefer (less than 24h) than after initial training, when HPC can support recall for several days. Although the mechanism underlying this faster disengagement is unknown, in our model we simulate the phenomenon by increasing the depotentiation probability in connections that are activated following memory retrieval: depotProb ij ← maxDepotProb t (11) where maxDepotProb is a tract-specific parameter that has a higher value in HPC tracts than in t neocortical tract. Subsequently, the depotentiation rate returns asymptotically to its base value: depotProb ← depotProb depotProbDecayRate − *( depotProb baseDepotProb − ) (12) where depotProbDecayRate is a global parameter. Hippocampal lesion. Hippocampal lesion is simulated by disconnecting the HPC region from the simulation. PSI infusion. Infusion of PSI into a region -- HPC or ACC -- is simulated by disabling potentiation and CI-AMPAR insertion for nine hours of simulation time, corresponding to the amount of time that the protein synthesis inhibitor anisomycin remains active in brain tissue (Wanisch & Wotjak, 2008). Inactivation. Reversible inactivation of HPC or ACC is modeled by transiently disabling activation of all units in the HPC or ACC region, respectively. 33 MODELING SYSTEMS RECONSOLIDATION Simulation environment The model is implemented as a C++ program and all simulations were executed on an Intel i5-2400 computer running the Debian Linux 8.4 operating system. Model fitting Simulations are controlled by the parameters indicated in the following tables. Parameter values were manually chosen to make simulation time courses approximate those reported in the referenced literature. Table 2: Global parameters Parameter actK minPsdSize maxPsdSize numSettleCycles trainNumStimCycles consNumStimCycles minInhib maxInhib inhibIncr Value 2.0 10 100 20 50 1 2.5 10 0.05 Description Steepness of the activation function Min value for a connection's psdSize Max value for a connection's psdSize Number of times the network is cycled for settling Number of times the learning rule is applied at training Number of times the learning rule is applied at consolidation Minimum inhibition level in a region Maximum inhibition level in a region Inhibition level adjustment rate Table 3: Region-specific parameters. HPC ACC SC0/1 Description 25 0.2 Number of units Target fraction of active units 25 0.2 25 0.2 Parameter numUnitsr kr 34 MODELING SYSTEMS RECONSOLIDATION Table 4: Tract-specific parameters. The "HPC" columns indicates values for the tracts that connect HPC with the other regions, and the "ACC" column for the tracts that connect ACC with SC0 and SC1. Name µt psdDecayRate01h cpAmparRemovalRate01h ciAmparInsertionRate01h ciAmparRemovalRate01h HPC ACC Description 0.004 Learning rate 0.08 0.01 0.01 0.1 0.1 2.0 2.0 0.015 0.015 CI-AMPAR removal rate in unpotentiated PSD shrink rate when unpopulated Constitutive removal rate for CP-AMPARs CI-AMPAR insertion rate in potentiated synapse baseDepotProb01h maxDepotProb01h depotProbDecayRate01h minNumCpAmpars minNumCiAmpars 0.002 0.0 0.0 0.05 0.03 0.03 0 0 2 2 synapse Probability of spontaneous depotentiation Prob. of depotentiation after reactivation depotProb decay rate towards baseDepotProb Num CP-AMPARs in an unpotentiated conn. Num CI-AMPARs in an unpotentiated conn. Results After training the network with a CS-US association, recall is tested by presenting the CS in the SC0 region, settling the system, and comparing the resulting activation pattern in the SC1 region with the trained US (a score of 1.0 indicates a perfect match). The descriptions below compare the recall scores in simulation runs where an intervention (lesion or inactivation) is performed with simulation runs without the intervention (labeled "baseline" in the diagrams). The score values in all diagrams are means of 100 simulation runs. Error bars indicate standard deviation. Consolidation and reconsolidation windows. HPC lesions produce memory deficits when performed in the consolidation or reconsolidation windows, but not otherwise. See Figure 3 and Figure 4. These simulations reproduce findings 5-7 and 10-11 of Table 1. 35 MODELING SYSTEMS RECONSOLIDATION Figure 3: Consolidation window. Simulated HPC lesions produce severe impairment when performed shortly after training, but not later. Recall tests are performed 7 days after lesioning, e.g. the data point corresponding to lesions performed 5 days after conditioning reflects recall tests executed on day 12. This simulates delays used in behavioral experiments to allow the animals to recover from surgery before testing. The baseline data points reflect recall tests at the corresponding time points, but without preceding lesions. 36 MODELING SYSTEMS RECONSOLIDATION Figure 4: Reconsolidation window: Simulated HPC lesions produce severe impairment when performed shortly after reactivation, but not later. Recall tests are performed 7 days after lesioning, i.e. the data points corresponding to lesions performed 0-23 hours after reactivation reflect recall tests executed on day 7 after reactivation, etc. The baseline data points correspond to recall tests at the corresponding time points, but without preceding lesions. 37 MODELING SYSTEMS RECONSOLIDATION Effect of systemic PSI infusion before training. Systemic PSI infusion before conditioning impairs formation of long-term memory but does not impair short-term memory, see Figure 5. This simulation reproduces finding 3 of Table 1. Figure 5: The effect of systemic PSI infusion before training. Short-term memory (1h) is unaffected, but long-term memory (24h) is severely impaired. 38 MODELING SYSTEMS RECONSOLIDATION Effect of PSI on consolidated memory. Systemic PSI infusion during maintenance, i.e. after completed systems consolidation, does not impair subsequent memory retrieval, see Figure 6. This simulation reproduces finding 4 of Table 1. Figure 6: Effect of PSI on consolidated memory. Systemic PSI infusion is simulated on day 30 after conditioning. The diagram shows recall performance before and after the intervention. PSI infusion in the maintenance phase does not affect recall performance. PSI infusion in HPC after reactivation. Post-reactivation PSI infusion in HPC does not cause immediate memory loss. Rather, the recall impairment develops over a period of more 39 MODELING SYSTEMS RECONSOLIDATION than 4 but less than 48 hours (Debiec et al., 2002), see Figure 7. This simulation reproduces findings 8 and 9 of Table 1. Figure 7: PSI infusion in HPC after reactivation. PSI is infused immediately after reactivation; recall tests are performed four hours later or 48 hours later. The impairment only manifests at the later time point. 40 MODELING SYSTEMS RECONSOLIDATION HPC/ACC-dependence for recall. Consolidation transforms memories from being HPC- dependent to being ACC-dependent for recall, see Figure 8. These simulations reproduce findings 1 and 2 of Table 1. Figure 8: HPC/ACC-dependence for recall. HPC inactivation impairs recall 3 days after training, but not at 30 days. ACC inactivation does not affect recall 3 days after training, but causes severe impairment at 30 days. Temporary ACC-independence after reactivation. Reactivation creates a transient HPC linkage which temporarily returns the memory to ACC-independence, see Figure 9. These simulations reproduce findings 12 and 13 of Table 1. 41 MODELING SYSTEMS RECONSOLIDATION Figure 9: Temporary ACC-independence after reactivation. Before reactivation ACC inactivation severely impairs recall of a consolidated memory. Six hours after reactivation neither ACC inactivation nor HPC activation produces significant impairment. At 24h after reactivation, ACC dependence has returned. Discussion We have presented an artificial neural network model of systems memory consolidation and reconsolidation that accounts for a broad range of findings from the empirical literature, including those from studies employing hippocampal lesions as well as ones using reversible inactivation of the hippocampus or anterior cingulate cortex. At the core of the model is a new connection design in which variable stability arises from simulation of receptor exchanges that have been observed in glutamatergic synapses. 42 MODELING SYSTEMS RECONSOLIDATION Reconsolidation It is worth noting that although the term "reconsolidation" suggests a recapitulation of consolidation, the model reflects our view that the two processes are quite different. Whereas systems consolidation is a gradual strengthening of the intra-neocortical connections that link together the components of a memory, systems reconsolidation consists in the restoration of stability in such synapses following reactivation-induced destabilization. HPC lesions in the "consolidation window" (the first couple of weeks after training) and in the "reconsolidation window" (the first day or two after reactivation) both result in memory deficits -- but for different reasons: HPC lesions in the consolidation window prematurely interrupts the hippocampal replay that drives establishment and strengthening of neocortical linkages, leaving a weak memory trace there. Loss of HPC in the reconsolidation window, in contrast, deprives neocortical neurons of the HPC stimulation that is needed to restabilize synapses that have been destabilized by reactivation-induced AMPAR exchange. Without such stimulation, the destabilized neocortical trace decays. Our model is the first to demonstrate results analogous to both systems consolidation and systems reconsolidation in an artificial neural network, and also to reproduce the effects of a number of experimental interventions described in the empirical literature. This was made possible by the use of a connection model with a plasticity mechanism inspired by biological synapses. Limitations Although the model is capable of reproducing a number of empirical results, it should be kept in mind that it is a restricted model with many simplifying assumptions. Thus, while the model only includes synaptic potentiation as a substrate for the creation of memory traces in 43 MODELING SYSTEMS RECONSOLIDATION systems consolidation and reconsolidation, it seems likely that other mechanisms are also at play, perhaps including axonal growth, synaptogenesis, and possibly neurogenesis (Frankland & Bontempi, 2005). Similarly, the simple receptor trafficking in the network connections is a crude model of synaptic potentiation, which is known to involve a large number of molecular processes, most of which are not well understood (Asok et al., 2019). We chose to model AMPAR trafficking because of the strong evidence of its importance for induction and maintenance of synaptic potentiation, as discussed in the introduction. Despite its limitations, it is interesting to find that a model as simple as this is able to capture a significant set of diverse findings from the consolidation/reconsolidation literature. Predictions In closing, we list a number of predictions that have been derived from the model and may be used to test its validity in future experiments. 1. In the model, post-reactivation susceptibility to HPC lesion is due to the CI-AMPAR to CP- AMPAR exchange that reactivation triggers in the consolidated connections in the ACC: subsequent CI-AMPAR restoration depends on HPC activity. A prediction that can be drawn from the model is therefore, that if reactivation is prevented from triggering AMPA receptor exchange in the ACC, then HPC lesion in the reconsolidation window will not impair recall. This could be tested by infusing a drug like GluA23Y into the ACC before reactivation. GluA23Y is a synthetic peptide that prevents endocytosis (removal) of CI-AMPARs from the synapse. In contrast, GluA23Y should not be able to prevent the amnestic effect of hippocampal lesion during the consolidation window, because in this case the impairment is not due to depotentiation but to interrupted consolidation. 44 MODELING SYSTEMS RECONSOLIDATION 2. The amnestic effect of post-reactivation PSI infusion into HPC is also abolished if destabilization of ACC connections is blocked. The reasoning is the same as above: If the ACC connections are not destabilized, then they do not need to be restabilized, and thus depriving them of hippocampal stimulation should have no amnestic effect. So the prediction is that GluA23Y infusion in ACC before reactivation would abolish the amnestic effect of post-reactivation PSI infusion in hippocampus. 3. Destabilization can also be prevented by selectively inhibiting GluN2B-containing NMDA receptors (Ben Mamou et al., 2006; Milton et al., 2013), cannabinoid receptor 1, or L-type voltage gated calcium channels (R. Kim et al., 2011). The model predicts that inhibiting any of these receptors or channels in ACC during reactivation would reduce or eliminate the amnestic effect of post-reactivation hippocampal lesion or PSI injection. 4. The model attributes the short duration of post-reactivation ACC-independence to accelerated depotentiation of linkage synapses in hippocampus. Blocking AMPAR endocytosis in hippocampus, e.g. by local GluA23Y infusion, should extend the durability of reactivation-induced hippocampal memory traces, and therefore lengthen the duration of post-reactivation ACC-independence. 5. In the model, post-reactivation hippocampal lesion or PSI infusion causes memory impairments by depriving neocortical linkage synapses of the stimulation they need to restabilize. If this is correct, then other means of blocking such stimulation should have the same amnestic effect. For example, we predict that prolonged (several days) reversible inactivation of hippocampus after reactivation will impair subsequent recall of the reactivated memory. 45 MODELING SYSTEMS RECONSOLIDATION 6. In the model, all that is needed for restabilization of a reactivated memory trace is repeated activation of its ACC linkage. This suggests that it may be possible to compensate for an inactivated hippocampus (as in prediction 5) by triggering reactivations externally, i.e. by reminders. This idea is related to a result reported by Lehmann et al. (2009), where repeated conditioning sessions were shown to significantly speed up the development of hippocampus-independence. Data availability statement All computer program files are available from the ModelDB database, accession number TBD (http://modeldb.yale.edu/TBD). References Abraham, W. C. (2003). How Long Will Long-Term Potentiation Last? Philosophical Transactions: Biological Sciences, 358(1432), 735 -- 744. Ackley, D., Hinton, G., & Sejnowski, T. (1985). A Learning Algorithm for Boltzmann Machines. Cognitive Science, 9(1), 147 -- 169. Alvarez, P., & Squire, L. R. (1994). Memory and the medial temporal-lobe: A simple network model. Proceedings of the National Academy of Sciences of the United States of America, 91(15), 7041 -- 7045. https://doi.org/10.1073/pnas.91.15.7041 Amaral, O. B., Osan, R., Roesler, R., & Tort, A. B. L. (2008). A synaptic reinforcement‐based model for transient amnesia following disruptions of memory consolidation and reconsolidation. Hippocampus, 18(6), 584 -- 601. https://doi.org/10.1002/hipo.20420 Anagnostaras, S. G., Maren, S., & Fanselow, M. S. (1999). Temporally graded retrograde amnesia of contextual fear after hippocampal damage in rats: Within-subjects examination. Journal of Neuroscience, 19(3), 1106 -- 1114. 46 MODELING SYSTEMS RECONSOLIDATION Asok, A., Leroy, F., Rayman, J. B., & Kandel, E. R. (2019). Molecular Mechanisms of the Memory Trace. Trends in Neurosciences, 42(1), 14 -- 22. https://doi.org/10.1016/j.tins.2018.10.005 Baldi, E., & Bucherelli, C. (2015). Brain sites involved in fear memory reconsolidation and extinction of rodents. Neuroscience and Biobehavioral Reviews, 53, 160 -- 190. https://doi.org/10.1016/j.neubiorev.2015.04.003 Baltaci, S. B., Mogulkoc, R., & Baltaci, A. K. (2019). Molecular Mechanisms of Early and Late LTP. Neurochemical Research, 44(2), 281 -- 296. https://doi.org/10.1007/s11064-018- 2695-4 Bartol, T. M., Bromer, C., Kinney, J. P., Chirillo, M. A., Bourne, J. N., Harris, K. M., & Sejnowski, T. J. (2015). Hippocampal spine head sizes are highly precise. BioRxiv, 016329. https://doi.org/10.1101/016329 Bayley, P. J., Gold, J. J., Hopkins, R. O., & Squire, L. R. (2005). The neuroanatomy of remote memory. Neuron, 46(5), 799 -- 810. https://doi.org/10.1016/j.neuron.2005.04.034 Ben Mamou, C., Gamache, K., & Nader, K. (2006). NMDA receptors are critical for unleashing consolidated auditory fear memories. Nature Neuroscience, 9(10), 1237 -- 1239. https://doi.org/10.1038/nn1778 Besnard, A., Caboche, J., & Laroche, S. (2012). Reconsolidation of memory: A decade of debate. Progress in Neurobiology, 99(1), 61 -- 80. https://doi.org/10.1016/j.pneurobio.2012.07.002 Bliss, T., & Collingridge, G. (1993). A Synaptic Model of Memory -- Long-Term Potentiation in the Hippocampus. Nature, 361(6407), 31 -- 39. https://doi.org/10.1038/361031a0 47 MODELING SYSTEMS RECONSOLIDATION Bliss, T., & Lømo, T. (1973). Long-Lasting Potentiation of Synaptic Transmission in Dentate Area of Anesthetized Rabbit Following Stimulation of Perforant Path. Journal of Physiology-London, 232(2), 331 -- 356. Bontempi, B., Laurent-Demir, C., Destrade, C., & Jaffard, R. (1999). Time-dependent reorganization of brain circuitry underlying long-term memory storage. Nature, 400(6745), 671 -- 675. https://doi.org/10.1038/23270 Bortolotto, Z., Bashir, Z., Davies, C., & Collingridge, G. (1994). A Molecular Switch Activated by Metabotropic Glutamate Receptors Regulates Induction of Long-Term Potentiation. Nature, 368(6473), 740 -- 743. https://doi.org/10.1038/368740a0 Bourtchouladze, R., Abel, T., Berman, N., Gordon, R., Lapidus, K., & Kandel, E. R. (1998). Different training procedures recruit either one or two critical periods for contextual memory consolidation, each of which requires protein synthesis and PKA. Learning & Memory, 5(4 -- 5), 365 -- 374. Cajal, S. R. Y. (1894). The Croonian Lecture: La Fine Structure des Centres Nerveux. Proceedings of the Royal Society of London, 55(331 -- 335), 444 -- 468. https://doi.org/10.1098/rspl.1894.0063 Canal, C. E., Chang, Q., & Gold, P. E. (2007). Amnesia produced by altered release of neurotransmitters after intraamygdala injections of a protein synthesis inhibitor. Proceedings of the National Academy of Sciences of the United States of America, 104(30), 12500 -- 12505. https://doi.org/10.1073/pnas.0705195104 Citri, A., & Malenka, R. C. (2008). Synaptic plasticity: Multiple forms, functions, and mechanisms. Neuropsychopharmacology, 33(1), 18 -- 41. https://doi.org/10.1038/sj.npp.1301559 48 MODELING SYSTEMS RECONSOLIDATION Clem, R. L., & Huganir, R. L. (2010). Calcium-Permeable AMPA Receptor Dynamics Mediate Fear Memory Erasure. Science, 330(6007), 1108 -- 1112. https://doi.org/10.1126/science.1195298 Clopath, C., Ziegler, L., Vasilaki, E., Buesing, L., & Gerstner, W. (2008). Tag-Trigger- Consolidation: A Model of Early and Late Long-Term-Potentiation and Depression. Plos Computational Biology, 4(12), e1000248. https://doi.org/10.1371/journal.pcbi.1000248 Cronholm, B., & Lagergren, A. (1959). Memory Disturbances After Electroconvulsive-Therapy. Acta Psychiatrica Et Neurologica, 34(3), 283 -- 310. https://doi.org/10.1111/j.1600- 0447.1959.tb07580.x Davis, H. P., & Squire, L. R. (1984). Protein synthesis and memory: A review. Psychological Bulletin, 96(3), 518 -- 559. de Oliveira Alvares, L., Einarsson, E. Ö., Santana, F., Crestani, A. P., Haubrich, J., Cassini, L. F., … Quillfeldt, J. A. (2012). Periodically reactivated context memory retains its precision and dependence on the hippocampus. Hippocampus, 22(5), 1092 -- 1095. https://doi.org/10.1002/hipo.20983 Debiec, J., LeDoux, J. E., & Nader, K. (2002). Cellular and systems reconsolidation in the hippocampus. Neuron, 36(3), 527 -- 538. Dede, A. J. O., & Smith, C. N. (2016). The Functional and Structural Neuroanatomy of Systems Consolidation for Autobiographical and Semantic Memory. In Current Topics in Behavioral Neurosciences. SpringerLink (pp. 1 -- 32). https://doi.org/10.1007/7854_2016_452 49 MODELING SYSTEMS RECONSOLIDATION Ding, H. K., Teixeira, C. M., & Frankland, P. W. (2008). Inactivation of the anterior cingulate cortex blocks expression of remote, but not recent, conditioned taste aversion memory. Learning & Memory, 15(5), 290 -- 293. https://doi.org/10.1101/lm.905008 Dudai, Y. (2004). The neurobiology of consolidations, or, how stable is the engram? Annual Review of Psychology, 55, 51 -- 86. https://doi.org/10.1146/annurev.psych.55.090902.142050 Duncan, C. (1949). The Retroactive Effect of Electroshock on Learning. Journal of Comparative and Physiological Psychology, 42(1), 32 -- 44. https://doi.org/10.1037/h0058173 Ecker, U. K. H., & Lewandowsky, S. (2012). Computational constraints in cognitive theories of forgetting. Frontiers in Cognition, 3, 400. https://doi.org/10.3389/fpsyg.2012.00400 Einarsson, E. Ö., & Nader, K. (2012). Involvement of the anterior cingulate cortex in formation, consolidation, and reconsolidation of recent and remote contextual fear memory. Learning & Memory, 19(10), 449 -- 452. https://doi.org/10.1101/lm.027227.112 Einarsson, E. Ö., Pors, J., & Nader, K. (2015). Systems Reconsolidation Reveals a Selective Role for the Anterior Cingulate Cortex in Generalized Contextual Fear Memory Expression. Neuropsychopharmacology, 40(2), 480 -- 487. https://doi.org/10.1038/npp.2014.197 Fanselow, M., & Kim, J. (1994). Acquisition of Contextual Pavlovian Fear Conditioning Is Blocked by Application of an Nmda Receptor Antagonist D,l-2-Amino-5- Phosphonovaleric Acid to the Basolateral Amygdala. Behavioral Neuroscience, 108(1), 210 -- 212. https://doi.org/10.1037//0735-7044.108.1.210 50 MODELING SYSTEMS RECONSOLIDATION Fonseca, R., Nägerl, U. V., & Bonhoeffer, T. (2006). Neuronal activity determines the protein synthesis dependence of long-term potentiation. Nature Neuroscience, 9(4), 478 -- 480. https://doi.org/10.1038/nn1667 Forcato, C., Burgos, V. L., Argibay, P. F., Molina, V. A., Pedreira, M. E., & Maldonado, H. (2007). Reconsolidation of declarative memory in humans. Learning & Memory, 14(4), 295 -- 303. https://doi.org/10.1101/lm.486107 Frankland, P. W., & Bontempi, B. (2005). The organization of recent and remote memories. Nature Reviews Neuroscience, 6(2), 119 -- 130. https://doi.org/10.1038/nrn1607 Frankland, P. W., Bontempi, B., Talton, L. E., Kaczmarek, L., & Silva, A. J. (2004). The involvement of the anterior cingulate cortex in remote contextual fear memory. Science, 304(5672), 881 -- 883. https://doi.org/10.1126/science.1094804 Frankland, P. W., Ding, H.-K., Takahashi, E., Suzuki, A., Kida, S., & Silva, A. J. (2006). Stability of recent and remote contextual fear memory. Learning & Memory, 13(4), 451 -- 457. https://doi.org/10.1101/lm.183406 Frey, U. (1997). Chapter 6 - Cellular Mechanisms Of Long-Term Potentiation: Late Maintenance. In J. W. Donahoe & V. Packard Dorsel (Eds.), Advances in Psychology (pp. 105 -- 128). https://doi.org/10.1016/S0166-4115(97)80092-1 Frey, U., Huang, Y., & Kandel, E. (1993). Effects of Camp Simulate a Late-Stage of Ltp in Hippocampal Ca1 Neurons. Science, 260(5114), 1661 -- 1664. https://doi.org/10.1126/science.8389057 Frey, U., Krug, M., Reymann, K. G., & Matthies, H. (1988). Anisomycin, an inhibitor of protein synthesis, blocks late phases of LTP phenomena in the hippocampal CA1 region in vitro. Brain Research, 452(1 -- 2), 57 -- 65. 51 MODELING SYSTEMS RECONSOLIDATION Gafford, G. M., Parsons, R. G., & Helmstetter, F. J. (2013). Memory accuracy predicts hippocampal mTOR pathway activation following retrieval of contextual fear memory. Hippocampus, 23(9), 842 -- 847. https://doi.org/10.1002/hipo.22140 Genzel, L., Rossato, J. I., Jacobse, J., Grieves, R. M., Spooner, P. A., Battaglia, F. P., … Morris, R. G. M. (2017). The Yin and Yang of Memory Consolidation: Hippocampal and Neocortical. Plos Biology, 15(1), e200531. https://doi.org/10.1371/journal.pbio.2000531 Gerard, R. (1955). Biological Roots of Psychiatry. Science, 122(3162), 225 -- 230. https://doi.org/10.1126/science.122.3162.225 Ghazal, P. (2016). AMPA receptor trafficking in recent vs. Remote memory. Brain Research, 1650, 203 -- 207. https://doi.org/10.1016/j.brainres.2016.09.010 Hardt, O., & Nadel, L. (2018). Systems consolidation revisited, but not revised: The promise and limits of optogenetics in the study of memory. Neuroscience Letters, 680, 54 -- 59. https://doi.org/10.1016/j.neulet.2017.11.062 Hardt, O., Nader, K., & Nadel, L. (2013). Decay happens: The role of active forgetting in memory. Trends in Cognitive Sciences, 17(3), 111 -- 120. https://doi.org/10.1016/j.tics.2013.01.001 Haubrich, J., Crestani, A. P., Cassini, L. F., Santana, F., Sierra, R. O., de Oliveira Alvares, L., & Quillfeldt, J. A. (2015). Reconsolidation Allows Fear Memory to Be Updated to a Less Aversive Level through the Incorporation of Appetitive Information. Neuropsychopharmacology, 40(2), 315 -- 326. https://doi.org/10.1038/npp.2014.174 Hebb, D. O. (1949). The Organization of Behavior: A Neuropsychological Approach. John Wiley & Sons. 52 MODELING SYSTEMS RECONSOLIDATION Helfer, P., & Shultz, T. R. (2018). Coupled feedback loops maintain synaptic long-term potentiation: A computational model of PKMzeta synthesis and AMPA receptor trafficking. PLOS Computational Biology, 14(5), e1006147. https://doi.org/10.1371/journal.pcbi.1006147 Henley, J. M., & Wilkinson, K. A. (2013). AMPA receptor trafficking and the mechanisms underlying synaptic plasticity and cognitive aging. Dialogues in Clinical Neuroscience, 15(1), 11 -- 27. Hong, I., Kim, J., Kim, J., Lee, S., Ko, H.-G., Nader, K., … Choi, S. (2013). AMPA receptor exchange underlies transient memory destabilization on retrieval. Proceedings of the National Academy of Sciences of the United States of America, 110(20), 8218 -- 8223. https://doi.org/10.1073/pnas.1305235110 Huang, Y., Li, X., & Kandel, E. (1994). Camp Contributes to Messy Fiber Ltp by Initiating Both a Covalently Mediated Early Phase and Macromolecular Synthesis-Dependent Late- Phase. Cell, 79(1), 69 -- 79. https://doi.org/10.1016/0092-8674(94)90401-4 Hupbach, A., Gomez, R., Hardt, O., & Nadel, L. (2007). Reconsolidation of episodic memories: A subtle reminder triggers integration of new information. Learning & Memory, 14(1 -- 2), 47 -- 53. https://doi.org/10.1101/lm.365707 Jalil, S. J., Sacktor, T. C., & Shouval, H. Z. (2015). Atypical PKCs in memory maintenance: The roles of feedback and redundancy. Learning & Memory (Cold Spring Harbor, N.Y.), 22(7), 344 -- 353. https://doi.org/10.1101/lm.038844.115 Jones, B. J., Pest, S. M., Vargas, I. M., Glisky, E. L., & Fellous, J.-M. (2015). Contextual reminders fail to trigger memory reconsolidation in aged rats and aged humans. Neurobiology of Learning and Memory, 120. https://doi.org/10.1016/j.nlm.2015.02.003 53 MODELING SYSTEMS RECONSOLIDATION Judge, M., & Quartermain, D. (1982). Characteristics of Retrograde-Amnesia Following Reactivation of Memory. Physiology & Behavior, 28(4), 585 -- 590. https://doi.org/10.1016/0031-9384(82)90034-8 Kandel, E. R. (2000). Cellular mechanisms of learning and the biological basis of individuality. In E. R. Kandel, J. H. Schwartz, & T. M. Jessel (Eds.), Principles of neural science (4th ed., pp. 1247 -- 1279). New York: McGraw Hill. Kandel, E. R., Dudai, Y., & Mayford, M. R. (2014). The Molecular and Systems Biology of Memory. Cell, 157(1), 163 -- 186. https://doi.org/10.1016/j.cell.2014.03.001 Kehlet, H., & Lunn, V. (1951). Retrograde Amnesia Following Electroshock Therapy. Nordisk Medicin, 45(19), 716 -- 719. Kessels, H. W., & Malinow, R. (2009). Synaptic AMPA Receptor Plasticity and Behavior. Neuron, 61(3), 340 -- 350. https://doi.org/10.1016/j.neuron.2009.01.015 Kim, J., Clark, R., & Thompson, R. (1995). Hippocampectomy Impairs the Memory of Recently, but Not Remotely, Acquired Trace Eyeblink Conditioned-Responses. Behavioral Neuroscience, 109(2), 195 -- 203. https://doi.org/10.1037/0735-7044.109.2.195 Kim, J., & Fanselow, M. (1992). Modality-specific retrograde-amnesia of fear. Science, 256(5057), 675 -- 677. https://doi.org/10.1126/science.1585183 Kim, R., Moki, R., & Kida, S. (2011). Molecular mechanisms for the destabilization and restabilization of reactivated spatial memory in the Morris water maze. Molecular Brain, 4, 9. https://doi.org/10.1186/1756-6606-4-9 Kitamura, T., Saitoh, Y., Takashima, N., Murayama, A., Niibori, Y., Ageta, H., … Inokuchi, K. (2009). Adult Neurogenesis Modulates the Hippocampus-Dependent Period of 54 MODELING SYSTEMS RECONSOLIDATION Associative Fear Memory. Cell, 139(4), 814 -- 827. https://doi.org/10.1016/j.cell.2009.10.020 Lah, S., & Miller, L. (2008). Effects of Temporal Lobe Lesions on Retrograde Memory: A Critical Review. Neuropsychology Review, 18(1), 24. https://doi.org/10.1007/s11065- 008-9053-2 Land, C., Bunsey, M., & Riccio, D. C. (2000). Anomalous properties of hippocampal lesion- induced retrograde amnesia. Psychobiology, 28(4), 476 -- 485. Langille, J. J., & Brown, R. E. (2018). The Synaptic Theory of Memory: A Historical Survey and Reconciliation of Recent Opposition. Frontiers in Systems Neuroscience, 12. https://doi.org/10.3389/fnsys.2018.00052 Lattal, K. M., & Abel, T. (2004). Behavioral impairments caused by injections of the protein synthesis inhibitor anisomycin after contextual retrieval reverse with time. Proceedings of the National Academy of Sciences of the United States of America, 101(13), 4667 -- 4672. https://doi.org/10.1073/pnas.0306546101 Lavenex, P., & Amaral, D. G. (2000). Hippocampal-neocortical interaction: A hierarchy of associativity. Hippocampus, 10(4), 420 -- 430. https://doi.org/10.1002/1098- 1063(2000)10:4<420::AID-HIPO8>3.0.CO;2-5 Lee, J. L. C. (2009). Reconsolidation: Maintaining memory relevance. Trends in Neurosciences, 32, 413 -- 420. https://doi.org/10.1016/j.tins.2009.05.002 Lee, J. L. C., Nader, K., & Schiller, D. (2017). An Update on Memory Reconsolidation Updating. Trends in Cognitive Sciences. https://doi.org/10.1016/j.tics.2017.04.006 55 MODELING SYSTEMS RECONSOLIDATION Lehmann, H., Sparks, F. T., Spanswick, S. C., Hadikin, C., McDonald, R. J., & Sutherland, R. J. (2009). Making context memories independent of the hippocampus. Learning & Memory, 16(7), 417 -- 420. https://doi.org/10.1101/lm.1385409 Lewis, D., Mahan, J., & Bregman, N. (1972). Cue-Dependent Amnesia in Rats. Journal of Comparative and Physiological Psychology, 81(2), 243-. https://doi.org/10.1037/h0033524 Ling, D. S. F., Benardo, L. S., Serrano, P. A., Blace, N., Kelly, M. T., Crary, John. F., & Sacktor, T. C. (2002). Protein kinase M zeta is necessary and sufficient for LTP maintenance. Nature Neuroscience, 5(4), 295 -- 296. https://doi.org/10.1038/nn829 Lisman, J. (2017). Criteria for identifying the molecular basis of the engram (CaMKII, PKMzeta). Molecular Brain, 10, 55. https://doi.org/10.1186/s13041-017-0337-4 Lynch, M. A. (2004). Long-term potentiation and memory. Physiological Reviews, 84(1), 87 -- 136. https://doi.org/10.1152/physrev.00014.2003 Malenka, R. C., & Bear, M. F. (2004). LTP and LTD: An embarrassment of riches. Neuron, 44(1), 5 -- 21. https://doi.org/10.1016/j.neuron.2004.09.012 Manns, J. R., Hopkins, R. O., & Squire, L. R. (2003). Semantic memory and the human hippocampus. Neuron, 38(1), 127 -- 133. https://doi.org/10.1016/S0896-6273(03)00146-6 Maren, S., Tocco, G., Standley, S., Baudry, M., & Thompson, R. (1993). Postsynaptic Factors in the Expression of Long-Term Potentiation (ltp) -- Increased Glutamate-Receptor Binding Following Ltp Induction in-Vivo. Proceedings of the National Academy of Sciences of the United States of America, 90(20), 9654 -- 9658. https://doi.org/10.1073/pnas.90.20.9654 56 MODELING SYSTEMS RECONSOLIDATION Maviel, T., Durkin, T. P., Menzaghi, F., & Bontempi, B. (2004). Sites of Neocortical Reorganization Critical for Remote Spatial Memory. Science, 305(5680), 96 -- 99. https://doi.org/10.1126/science.1098180 McClelland, J. L., McNaughton, B. L., & O'Reilly, R. C. (1995). Why there are complementary learning systems in the hippocampus and neocortex: Insights from the successes and failures of connectionist models of learning and memory. Psychological Review, 102, 419 -- 457. https://doi.org/10.1037/0033-295X.102.3.419 McCormack, S. G., Stornetta, R. L., & Zhu, J. J. (2006). Synaptic AMPA receptor exchange maintains bidirectional plasticity. Neuron, 50(1), 75 -- 88. https://doi.org/10.1016/j.neuron.2006.02.027 Meeter, M., & Murre, J. M. J. (2005). Tracelink: A model of consolidation and amnesia. Cognitive Neuropsychology, 22(5), 559 -- 587. https://doi.org/10.1080/02643290442000194 Migues, P. V., Hardt, O., Wu, D. C., Gamache, K., Sacktor, T. C., Wang, Y. T., & Nader, K. (2010). PKMzeta maintains memories by regulating GluR2-dependent AMPA receptor trafficking. Nature Neuroscience, 13(5), 630 -- 634. https://doi.org/10.1038/nn.2531 Migues, P. V., Liu, L., Archbold, G. E. B., Einarsson, E. Ö., Wong, J., Bonasia, K., … Hardt, O. (2016). Blocking Synaptic Removal of GluA2-Containing AMPA Receptors Prevents the Natural Forgetting of Long-Term Memories. Journal of Neuroscience, 36(12), 3481 -- 3494. https://doi.org/10.1523/JNEUROSCI.3333-15.2016 Milekic, M. H., & Alberini, C. M. (2002). Temporally graded requirement for protein synthesis following memory reactivation. Neuron, 36(3), 521 -- 525. https://doi.org/10.1016/S0896- 6273(02)00976-5 57 MODELING SYSTEMS RECONSOLIDATION Milner, P. (1989). A Cell Assembly Theory of Hippocampal Amnesia. Neuropsychologia, 27(1), 23 -- 30. https://doi.org/10.1016/0028-3932(89)90087-0 Milton, A. L., Merlo, E., Ratano, P., Gregory, B. L., Dumbreck, J. K., & Everitt, B. J. (2013). Double Dissociation of the Requirement for GluN2B- and GluN2A- Containing NMDA Receptors in the Destabilization and Restabilization of a Reconsolidating Memory. Journal of Neuroscience, 33(3), 1109 -- 1115. https://doi.org/10.1523/JNEUROSCI.3273- 12.2013 Misanin, J. R., Miller, R. R., & Lewis, D. J. (1968). Retrograde Amnesia Produced by Electroconvulsive Shock after Reactivation of a Consolidated Memory Trace. Science, 160(3827), 554 -- 555. https://doi.org/10.1126/science.160.3827.554 Miserendino, M. J. D., Sananes, C. B., Melia, K. R., & Davis, M. (1990). Blocking of acquisition but not expression of conditioned fear-potentiated startle by NMDA antagonists in the amygdala. Nature, 345(6277), 716. https://doi.org/10.1038/345716a0 Monfils, M.-H., Cowansage, K. K., Klann, E., & LeDoux, J. E. (2009). Extinction- Reconsolidation Boundaries: Key to Persistent Attenuation of Fear Memories. Science, 324(5929), 951 -- 955. https://doi.org/10.1126/science.1167975 Moscovitch, M., Cabeza, R., Winocur, G., & Nadel, L. (2016). Episodic Memory and Beyond: The Hippocampus and Neocortex in Transformation. Annual Review of Psychology, 67(1), 105 -- 134. https://doi.org/10.1146/annurev-psych-113011-143733 Murre, J. M. J. (1996). TraceLink: A model of amnesia and consolidation of memory. Hippocampus, 6(6), 675 -- 684. 58 MODELING SYSTEMS RECONSOLIDATION Nabavi, S., Fox, R., Proulx, C. D., Lin, J. Y., Tsien, R. Y., & Malinow, R. (2014). Engineering a memory with LTD and LTP. Nature, 511(7509), 348-+. https://doi.org/10.1038/nature13294 Nadel, L., & Hardt, O. (2010). Update on memory systems and processes. Neuropsychopharmacology, 36(1), 251 -- 273. Nadel, L., & Moscovitch, M. (1997). Memory consolidation, retrograde amnesia and the hippocampal complex. Current Opinion in Neurobiology, 7(2), 217 -- 227. https://doi.org/10.1016/S0959-4388(97)80010-4 Nadel, L., Winocur, G., Ryan, L., & Moscovitch, M. (2007). Systems consolidation and hippocampus: Two views. Debates in Neuroscience, 1(2 -- 4), 55 -- 66. https://doi.org/10.1007/s11559-007-9003-9 Nader, K., & Einarsson, E. Ö. (2010). Memory reconsolidation: An update. Annals of the New York Academy of Sciences, 1191, 27 -- 41. https://doi.org/10.1111/j.1749- 6632.2010.05443.x Nader, K., Schafe, G. E., & Le Doux, J. E. (2000). Fear memories require protein synthesis in the amygdala for reconsolidation after retrieval. Nature, 406(6797), 722 -- 726. https://doi.org/10.1038/35021052 Okubo-Suzuki, R., Saitoh, Y., Shehata, M., Zhao, Q., Enomoto, H., & Inokuchi, K. (2016). Frequency-specific stimulations induce reconsolidation of long-term potentiation in freely moving rats. Molecular Brain, 9, 36. https://doi.org/10.1186/s13041-016-0216-4 Pastalkova, E., Serrano, P., Pinkhasova, D., Wallace, E., Fenton, A. A., & Sacktor, T. C. (2006). Storage of spatial information by the maintenance mechanism of LTP. Science, 313(5790), 1141 -- 1144. https://doi.org/10.1126/science.1128657 59 MODELING SYSTEMS RECONSOLIDATION Penfield, W., & Milner, B. (1958). Memory Deficit Produced by Bilateral Lesions in the Hippocampal Zone. A.M.A. Archives of Neurology & Psychiatry, 79(5), 475 -- 497. https://doi.org/10.1001/archneurpsyc.1958.02340050003001 Plant, K., Pelkey, K. A., Bortolotto, Z. A., Morita, D., Terashima, A., McBain, C. J., … Isaac, J. T. R. (2006). Transient incorporation of native GluR2-lacking AMPA receptors during hippocampal long-term potentiation. Nature Neuroscience, 9(5), 602 -- 604. https://doi.org/10.1038/nn1678 Platt, S. R. (2007). The role of glutamate in central nervous system health and disease -- A review. The Veterinary Journal, 173(2), 278 -- 286. https://doi.org/10.1016/j.tvjl.2005.11.007 Przybyslawski, J., & Sara, S. J. (1997). Reconsolidation of memory after its reactivation. Behavioural Brain Research, 84(1 -- 2), 241 -- 246. https://doi.org/10.1016/S0166- 4328(96)00153-2 Quirk, G. J., & Mueller, D. (2008). Neural mechanisms of extinction learning and retrieval. Neuropsychopharmacology, 33(1), 56 -- 72. https://doi.org/10.1038/sj.npp.1301555 Rao-Ruiz, P., Rotaru, D. C., van der Loo, R. J., Mansvelder, H. D., Stiedl, O., Smit, A. B., & Spijker, S. (2011). Retrieval-specific endocytosis of GluA2-AMPARs underlies adaptive reconsolidation of contextual fear. Nature Neuroscience, 14(10), 1302-U117. https://doi.org/10.1038/nn.2907 Rossato, J. I., Bevilaqua, L. R. M., Medina, J. H., Izquierdo, I., & Cammarota, M. (2006). Retrieval induces hippocampal-dependent reconsolidation of spatial memory. Learning & Memory, 13(4), 431 -- 440. https://doi.org/10.1101/lm.315206 60 MODELING SYSTEMS RECONSOLIDATION Rudy, J. W., Biedenkapp, J. C., & O'Reilly, R. C. (2005). Prefrontal cortex and the organization of recent and remote memories: An alternative view. Learning & Memory, 12(5), 445 -- 446. https://doi.org/10.1101/lm.97905 Runyan, J. D., Moore, A. N., & Dash, P. K. (2019). Coordinating what we've learned about memory consolidation: Revisiting a unified theory. Neuroscience & Biobehavioral Reviews, 100, 77 -- 84. https://doi.org/10.1016/j.neubiorev.2019.02.010 Sachser, R. M., Santana, F., Crestani, A. P., Lunardi, P., Pedraza, L. K., Quillfeldt, J. A., … de Oliveira Alvares, L. (2016). Forgetting of long-term memory requires activation of NMDA receptors, L-type voltage-dependent Ca2+ channels, and calcineurin. Scientific Reports, 6, 22771. https://doi.org/10.1038/srep22771 Sacktor, T. C. (2011). How does PKMζ maintain long-term memory? Nature Reviews Neuroscience, 12(1), 9 -- 15. https://doi.org/10.1038/nrn2949 Schiller, D., & Phelps, E. A. (2011). Does reconsolidation occur in humans? Frontiers in Behavioral Neuroscience, 5. https://doi.org/10.3389/fnbeh.2011.00024 Schneider, A. M., & Sherman, W. (1968). Amnesia: A function of the temporal relation of footshock to electroconvulsive shock. Science (New York, N.Y.), 159(3811), 219 -- 221. https://doi.org/10.1126/science.159.3811.219 Scoville, W., & Milner, B. (1957). Loss of Recent Memory After Bilateral Hippocampal Lesions. Journal of Neurology Neurosurgery and Psychiatry, 20(1), 11 -- 21. https://doi.org/10.1136/jnnp.20.1.11 Sederberg, P. B., Gershman, S. J., Polyn, S. M., & Norman, K. A. (2011). Human memory reconsolidation can be explained using the temporal context model. Psychonomic Bulletin & Review, 18, 455 -- 468. https://doi.org/10.3758/s13423-011-0086-9 61 MODELING SYSTEMS RECONSOLIDATION Serrano, P., Yao, Y. D., & Sacktor, T. C. (2005). Persistent phosphorylation by protein kinase M zeta maintains late-phase long-term potentiation. Journal of Neuroscience, 25(8), 1979 -- 1984. https://doi.org/10.1523/JNEUROSCI.5132-04.2005 Shema, R., Hazvi, S., Sacktor, T. C., & Dudai, Y. (2009). Boundary conditions for the maintenance of memory by PKM zeta in neocortex. Learning & Memory, 16(2), 122 -- 128. https://doi.org/10.1101/lm.1183309 Shema, R., Sacktor, T. C., & Dudai, Y. (2007). Rapid erasure of long-term memory associations in the cortex by an inhibitor of PKM zeta. Science, 317(5840), 951 -- 953. https://doi.org/10.1126/science.1144334 Shultz, T. R. (2003). Computational developmental psychology. MIT Press. Sierra, R. O., Pedraza, L. K., Zanona, Q. K., Santana, F., Boos, F. Z., Crestani, A. P., … Quillfeldt, J. A. (2017). Reconsolidation-induced Rescue of a Remote Fear Memory Blocked by an Early Cortical Inhibition: Involvement of the Anterior Cingulate Cortex and the Mediation by the Thalamic Nucleus Reuniens. Hippocampus, 27(5), 596 -- 607. https://doi.org/10.1002/hipo.22715 Smolen, P., Baxter, D. A., & Byrne, J. H. (2012). Molecular Constraints on Synaptic Tagging and Maintenance of Long-Term Potentiation: A Predictive Model. Plos Computational Biology, 8(8), e1002620. https://doi.org/10.1371/journal.pcbi.1002620 Sossin, W. S. (2008). Molecular memory traces. In W. S. Sossin, J. C. Lacaille, V. F. Castellucci, & S. Belleville (Eds.), Essence of Memory (Vol. 169, pp. 3 -- 25). Amsterdam: Elsevier Science Bv. Spear, N. E. (1973). Retrieval of memory in animals. Psychological Review, 80(3), 163 -- 194. https://doi.org/10.1037/h0034326 62 MODELING SYSTEMS RECONSOLIDATION Squire, L. R., & Alvarez, P. (1995). Retrograde amnesia and memory consolidation: A neurobiological perspective. Current Opinion in Neurobiology, 5(2), 169 -- 177. https://doi.org/10.1016/0959-4388(95)80023-9 Squire, L. R., & Cohen, N. (1979). Memory and amnesia: Resistance to disruption develops for years after learning. Behavioral and Neural Biology, 25(1), 115 -- 125. https://doi.org/10.1016/S0163-1047(79)90841-0 Squire, L. R., Genzel, L., Wixted, J. T., & Morris, R. G. (2015). Memory Consolidation. Cold Spring Harbor Perspectives in Biology, 7(8), a021766. https://doi.org/10.1101/cshperspect.a021766 Squire, L. R., & Zola-Morgan, S. (1991). The Medial Temporal-Lobe Memory System. Science, 253(5026), 1380 -- 1386. https://doi.org/10.1126/science.1896849 Stevens, C. F. (1998). A million dollar question: Does LTP = memory? Neuron, 20(1), 1 -- 2. https://doi.org/10.1016/S0896-6273(00)80426-2 Sutherland, R. J., Sparks, F. T., & Lehmann, H. (2010). Hippocampus and retrograde amnesia in the rat model: A modest proposal for the situation of systems consolidation. Neuropsychologia, 48(8), 2357 -- 2369. https://doi.org/10.1016/j.neuropsychologia.2010.04.015 Takehara, K., Kawahara, S., & Kirino, Y. (2003). Time-dependent reorganization of the brain components underlying memory retention in trace eyeblink conditioning. Journal of Neuroscience, 23(30), 9897 -- 9905. Teyler, T. J., & Discenna, P. (1986). The Hippocampal Memory Indexing Theory. Behavioral Neuroscience, 100(2), 147 -- 154. https://doi.org/10.1037/0735-7044.100.2.147 63 MODELING SYSTEMS RECONSOLIDATION Teyler, T. J., & Rudy, J. W. (2007). The hippocampal indexing theory and episodic memory: Updating the index. Hippocampus, 17(12), 1158 -- 1169. https://doi.org/10.1002/hipo.20350 Thompson, R., & Dean, W. (1955). A Further Study on the Retroactive Effect of Ecs. Journal of Comparative and Physiological Psychology, 48(6), 488 -- 491. https://doi.org/10.1037/h0041777 Walker, D. L., & Davis, M. (2002). The role of amygdala glutamate receptors in fear learning, fear-potentiated startle, and extinction. Pharmacology Biochemistry and Behavior, 71(3), 379 -- 392. https://doi.org/10.1016/S0091-3057(01)00698-0 Walker, M. P., Brakefield, T., Hobson, J. A., & Stickgold, R. (2003). Dissociable stages of human memory consolidation and reconsolidation. Nature, 425(6958), 616 -- 620. https://doi.org/10.1038/nature01930 Wang, S.-H., & Morris, R. G. M. (2010). Hippocampal-neocortical interactions in memory formation, consolidation, and reconsolidation. Annual Review of Psychology, 61, 49 -- 79, C1-4. https://doi.org/10.1146/annurev.psych.093008.100523 Wanisch, K., & Wotjak, C. T. (2008). Time course and efficiency of protein synthesis inhibition following intracerebral and systemic anisomycin treatment. Neurobiology of Learning and Memory, 90(3), 485 -- 494. https://doi.org/10.1016/j.nlm.2008.02.007 Westmark, P. R., Westmark, C. J., Wang, S., Levenson, J., O'Riordan, K. J., Burger, C., & Malter, J. S. (2010). Pin1 and PKM zeta Sequentially Control Dendritic Protein Synthesis. Science Signaling, 3(112), ra18. https://doi.org/10.1126/scisignal.2000451 64 MODELING SYSTEMS RECONSOLIDATION Wiltgen, B. J., Brown, R. A. M., Talton, L. E., & Silva, A. J. (2004). New Circuits for Old Memories: The Role of the Neocortex in Consolidation. Neuron, 44(1), 101 -- 108. https://doi.org/10.1016/j.neuron.2004.09.015 Wiltgen, B. J., Zhou, M., Cai, Y., Balaji, J., Karlsson, M. G., Parivash, S. N., … Silvaz, A. J. (2010). The hippocampus plays a selective role in the retrieval of detailed contextual memories. Current Biology, 20(15), 1336 -- 1344. https://doi.org/10.1016/j.cub.2010.06.068 Winocur, G. (1990). Anterograde and retrograde amnesia in rats with dorsal hippocampal or dorsomedial thalamic lesions. Behavioural Brain Research, 38(2), 145 -- 154. https://doi.org/10.1016/0166-4328(90)90012-4 Winocur, G., Frankland, P. W., Sekeres, M., Fogel, S., & Moscovitch, M. (2009). Changes in context-specificity during memory reconsolidation: Selective effects of hippocampal lesions. Learning & Memory, 16(11), 722 -- 729. https://doi.org/10.1101/lm.1447209 Winocur, G., Moscovitch, M., & Bontempi, B. (2010). Memory formation and long-term retention in humans and animals: Convergence towards a transformation account of hippocampal-neocortical interactions. Neuropsychologia, 48(8), 2339 -- 2356. https://doi.org/10.1016/j.neuropsychologia.2010.04.016 Winocur, G., Moscovitch, M., & Sekeres, M. J. (2013). Factors affecting graded and ungraded memory loss following hippocampal lesions. Neurobiology of Learning and Memory, 106, 351 -- 364. https://doi.org/10.1016/j.nlm.2013.10.001 Winocur, G., Sekeres, M. J., Binns, M. A., & Moscovitch, M. (2013). Hippocampal lesions produce both nongraded and temporally graded retrograde amnesia in the same rat. Hippocampus, 23(5), 330 -- 341. https://doi.org/10.1002/hipo.22093 65 MODELING SYSTEMS RECONSOLIDATION Wittenberg, G., & Tsien, J. (2002). An emerging molecular and cellular framework for memory processing by the hippocampus. Trends in Neurosciences, 25(10), 501 -- 505. https://doi.org/10.1016/S0166-2236(02)02231-2 Zhang, Y., Smolen, P., Baxter, D. A., & Byrne, J. H. (2010). The sensitivity of memory consolidation and reconsolidation to inhibitors of protein synthesis and kinases: Computational analysis. Learning & Memory, 17(9), 428 -- 439. https://doi.org/10.1101/lm.1844010 Zola-Morgan, S. M., & Squire, L. R. (1990). The primate hippocampal formation: Evidence for a time-limited role in memory storage. Science (New York, N.Y.), 250(4978), 288 -- 290. https://doi.org/10.1126/science.2218534 66
1608.04935
1
1608
2016-08-17T12:07:14
Identification of interictal epileptic networks from dense-EEG
[ "q-bio.NC" ]
Epilepsy is a network disease. The epileptic network usually involves spatially distributed brain regions. In this context, noninvasive M/EEG source connectivity is an emerging technique to identify functional brain networks at cortical level from noninvasive recordings. In this paper, we analyze the effect of the two key factors involved in EEG source connectivity processing: i) the algorithm used in the solution of the EEG inverse problem and ii) the method used in the estimation of the functional connectivity. We evaluate four inverse solutions algorithms and four connectivity measures on data simulated from a combined biophysical/physiological model to generate realistic interictal epileptic spikes reflected in scalp EEG. We use a new network-based similarity index (SI) to compare between the network identified by each of the inverse/connectivity combination and the original network generated in the model. The method will be also applied on real data recorded from one epileptic patient who underwent a full presurgical evaluation for drug-resistant focal epilepsy. In simulated data, results revealed that the selection of the inverse/connectivity combination has a significant impact on the identified networks. Results suggested that nonlinear methods for measuring the connectivity are more efficient than the linear one. The wMNE inverse solution showed higher performance than dSPM, cMEM and sLORETA. In real data, the combination (wMNE/PLV) led to a very good matching between the interictal epileptic network identified from noninvasive EEG recordings and the network obtained from connectivity analysis of intracerebral EEG recordings. These results suggest that source connectivity method, when appropriately configured, is able to extract highly relevant diagnostic information about networks involved in interictal epileptic spikes from non-invasive dense-EEG data.
q-bio.NC
q-bio
Identification ofinterictal epileptic networks from dense- EEG Mahmoud Hassan1,2, Isabelle Merlet1,2, Ahmad Mheich1,2,4, Aya Kabbara1,2,4, Arnaud Biraben1,2,3, Anca Nica3 and Fabrice Wendling1,2 1 INSERM, U1099, Rennes, F-35000, France 2 Université de Rennes 1, LTSI, F-35000, France 3 Neurology department, CHU, Rennes, France 4AZM center-EDST, Lebanese University, Tripoli, Lebanon Corresponding author: Mahmoud Hassan [email protected] Tel: +33 2 23 23 56 05 / 62 20 Fax: +33 2 23 23 69 17 1 Abstract Epilepsy is a network disease. The epileptic network usually involves spatially distributed brain regions. In this context, noninvasive M/EEG source connectivity is an emerging technique to identify functional brain networks at cortical level from noninvasive recordings. In this paper, we analyze the effect of the two key factors involved in EEG source connectivity processing: i) the algorithm used in the solution of the EEG inverse problem and ii) the method used in the estimation of the functional connectivity. We evaluate four inverse solutions algorithms (dSPM, wMNE, sLORETA and cMEM) and four connectivity measures (r2, h2, PLV, and MI) on data simulated from a combined biophysical/physiological model to generate realistic interictal epileptic spikes reflected in scalp EEG. We use a new network-based similarity index (SI) to compare between the network identified by each of the inverse/connectivity combination and the original network generated in the model. The method will be also applied on real data recorded from one epileptic patient who underwent a full presurgical evaluation for drug-resistant focal epilepsy. In simulated data, results revealed that the selection of the inverse/connectivity combination has a significant impact on the identified networks. Results suggested that nonlinear methods (nonlinear correlation coefficient, phase synchronization and mutual information) for measuring the connectivity are more efficient than the linear one (the cross correlation coefficient). The wMNE inverse solution showed higher performance than dSPM, cMEM and sLORETA. In real data, the combination (wMNE/PLV) led to a very good matching between the interictal epileptic network identified from noninvasive EEG recordings and the network obtained from connectivity analysis of intracerebral EEG recordings. These results suggest that source connectivity method, when appropriately configured, is able to extract highly relevant diagnostic information about networks involved in interictal epileptic spikes from non-invasive dense-EEG data. 2 Introduction Epilepsy is a network disease (Engel Jr et al. 2013). Over the two past decades, the concept of "epileptic focus" has progressively evolved toward that of "epileptic network" (Kramer and Cash 2012; Laufs 2012). Indeed, with the progress of functional neuroimaging techniques, many studies confirmed that the epileptic zone (EZ) can rarely be reduced to a circumscribed brain area (Bartolomei et al. 2001) as it very often involves distinct brain regions generating both interictal (Bourien et al. 2005) and ictal activity (Bourien et al. 2004). Among the investigation techniques classically used in the diagnostic of epilepsy, electrophysiological recordings (typically, magneto- and electro-encephalography including depth- EEG, referred to as M/EEG) are still extensively used to localize and delineate the EZ in a patient-specific context. Regarding the numerous methods proposed to process the recorded data; those aimed at characterizing brain connectivity are particularly suitable to identify networks involved during epileptiform activity (both interictal and ictal). In the context of invasive EEG signals (intracranial EEG, stereo-EEG and electrocorticoGraphy –EcoG-) recorded in patients candidate to surgery, these "connectivity" methods have been a topic of extensive research (see (van Mierlo et al. 2014) for recent review). For instance, the coherence function was shown to localize the seizure onset (Gotman 1987), similarity indexes were used to distinguish a preictal state from the ongoing interictal activity (Le Van Quyen et al. 2005; Mormann et al. 2000). Nonlinear regression analysis was applied to intracerebral signals to characterize connectivity patterns at the seizure onset (Bourien et al. 2004). Readers may refer to previous reviews for more detailed information about brain connectivity methods applied to non-invasive (van Mierlo et al. 2014) and invasive EEG signals (Wendling et al. 2010) in drug-resistant focal epilepsies. In the context of scalp M/EEG recording, connectivity methods have received less attention as compared with invasive EEG. A number of studies performed at the level of electrodes and focused on ictal periods have been reported aiming at analyzing seizure propagation (Gotman 1983) or to determine the seizure onset side (Caparos et al. 2006), for instance. For interictal periods, few connectivity studies made use of dense EEG and phase synchronization (Ramon and Holmes 2013) to identify epileptic sites. One reason for this paucity of studies may lie in the intricate interpretation of connectivity measures obtained from scalp recordings. Indeed, this interpretation is not straightforward as signals are severely corrupted by the effects of volume conduction (Schoffelen and Gross 2009). Interestingly, some recent studies showed how to overcome this limitation. In line with previous cognitive studies (Astolfi et al. 2007; Babiloni et al. 2005; Betti et al. 2013; Bola and Sabel 2015; David et al. 2003; David et al. 2002; de Pasquale et al. 2010; Hassan et al. 2015a; Hassan et al. 2014; Hassan and Wendling 2015; Hipp et al. 2011; Hoechstetter et al. 2004; Liljeström et al. 2015; Schoffelen and Gross 2009), the basic principle is to estimate functional connectivity at the level of 3 brain sources reconstructed from scalp signals. These methods, referred to as "source connectivity" were applied to both interictal EEG (Coito et al. 2015; Song et al. 2013; Vecchio et al. 2014) and MEG signals (Dai et al. 2012; Malinowska et al. 2014) as well as to EEG signals recorded during seizures (Ding et al. 2007; Jiruska et al. 2013; Lu et al. 2012) or resting states (Adebimpe et al. 2016; Coito et al. 2016). Although these approaches all include two steps (M/EEG inverse problem followed by source connectivity estimation), they strongly differ from a methodological viewpoint. Indeed, various algorithms were used to reconstruct brain sources and both functional and effective connectivity measures were utilized to assess statistical couplings among time series associated with reconstructed sources. Therefore, a central issue is the impact of selected methods (EEG inverse solution and connectivity measure) on the topological/statistical properties of identified epileptic networks activated during paroxysmal activity. In this paper, we report a quantitative comparison of methods aimed at identifying cortical epileptic networks from scalp EEG data. The novelty of this work is twofold. First, our comparative study includes simulated dense EEGs generated from physiologically- and biophysically-plausible models of distributed and coupled epileptic sources. To our knowledge, no previous study has reported results on the performance of source connectivity methods based on a "ground truth" provided by realistic computational models of interictal EEG signals (recorded later in time than the dense EEG recordings). Second, in line with a recent analysis performed on MEG data (Malinowska et al. 2014), networks estimated from real scalp dense EEG are compared with those obtained from depth-EEG recordings (SEEG). Materials and Methods A. Inverse solution algorithms Given the equivalent current dipole model, EEG signals X(t) recorded from M channels can be considered as linear combinations of P time-varying current dipole sources S(t): where G[M, P] is the lead field matrix and N(t) is the noise. As G is known, the EEG inverse problem consists of estimating the unknown sources from X(t). Several algorithms have been proposed to solve this problem based on different assumptions about spatial and temporal properties of sources and regularization constraints. Here, we chose to evaluate the four algorithms implemented in Brainstorm (Tadel et al. 2011). 1) Weighted Minimum Norm Estimate (wMNE) Minimum norm estimates (MNE) originally proposed by (Hämäläinen and Ilmoniemi 1994) are based on a search for the solution with minimum power using the L2 norm to regularize the problem. A common expression for MNE resolution 4 X(t)GS(t)N(t)S(t) matrix is where is the regularization parameter and C represents the noise covariance matrix. The weighted MNE (wMNE) algorithm compensates for the tendency of MNE to favor weak and surface sources (Hämäläinen 2005). This is achieved by introducing a weighting matrix WX in that adjusts the properties of the solution by reducing the bias inherent to the standard MNE solution. Classically, W X is a diagonal matrix built from matrix G with non-zero terms inversely proportional to the norm of the lead field vectors. 2) dynamical Statistical Parametric Mapping (dSPM) The dSPM is based on the MNE solution (Dale et al. 2000). For dSPM, the normalization matrix contains the minimum norm estimates of the noise at each source (Caparos et al. 2006), derived from the noise covariance matrix, defined as: Where . 3) Standardized low resolution brain electromagnetic tomography (sLORETA) sLORETA uses the source distribution estimated from MNE and standardizes it a posteriori by the variance of each estimated dipole source: where . The sLORETA inverse method has been originally described using the whole brain volume as source space (Pascual-Marqui 2002). For the present study, in order to ease the comparison with other methods, we restricted the source space to the neocortical surface. 4) Standard Maximum Entropy on the Mean (cMEM) The Maximum Entropy on the Mean (MEM) solver is based on a probabilistic method where inference on the current source intensities is estimated from the data, which is the basic idea of the maximum of entropy. The first application of MEM to 5 TT1MNEGGGCGS()T1TwMNEXXGWGCGWXS()dSPMdSPMMNESWS2TdSPMMNEMNEWdiagSCS()sLORETAsLORETAMNESWS2TTsLORETAMNEMNEMNEWdiagSGdiagSGGCS()(()) electromagnetic source localization was reported in (Clarke and Janday 1989). The main feature of this method is its ability to recover the spatial extent of the underlying sources. Its solution is assessed by finding the closest distribution of source intensities to a reference distribution in which source intensities are organized into cortical parcels showing homogeneous activation state (parallel cortical macro-columns with synchronized activity). In addition a constraint of local spatial smoothness in each parcel can be introduced (Chowdhury et al. 2013). B. Connectivity measures We selected four methods that have been widely used to estimate functional brain connectivity from electrophysiological signals (local field potentials, depth-EEG or EEG/MEG) (see (Wendling et al. 2009) for review). These measures were chosen to cover the main families of connectivity methods (linear and nonlinear regression, phase synchronization and mutual information). Briefly, concerning the regression approaches, the linear cross-correlation coefficient is only limited to the detection of the linear properties of the relationships between time series. However, mechanisms at the origin of EEG signals were shown to have strong nonlinear behaviors (Pereda et al. 2005). Thus, we have selected three nonlinear connectivity measures. The nonlinear regression where the basic idea is to evaluate the dependency of two signals from signal samples only and independently of the type of relation between the two signals. Concerning the phase synchronization measure, the method estimates the instantaneous phase of each signal and then computes a quantity based on co-variation of extracted phases to determine the degree of relationship. Finally, the mutual information method is based on the probability and information theory to measures mutual dependence between two variables. More technical details about the four methods are presented hereafter: 1) Cross-correlation coefficient (r2) The cross-correlation coefficient measures the linear correlation between two variables x and y as a function of their time delay ( ). Referred to as the linear correlation coefficient, it is defined as: where and cov denote the standard deviation and the covariance, respectively. 2) Nonlinear correlation coefficient (h2) The nonlinear correlation coefficient (h2) is a non-parametric measure of the nonlinear relationship between two time series x and y. In practice, the nonlinear relation between the two time series is approximated by a piecewise linear curve. 6 222()()cov((),())max()xtytxyxtytr where and f(x) is the linear piecewise approximation of the nonlinear regression curve. 3) Mutual information (MI) The mutual information (MI) between signal x and y is defined as: where is the joint probability of x=xi and y=yj. In the case of no relationship between x and y, = , so that the MI is zero for independent processes. Otherwise, will be positive, attaining its maximal value for identical signals. 4) Phase Locking Value (PLV) For two signals x and y, the phase locking value is defined as: where and are the unwrapped phases of the signals x and y at time t. The denotes the average over time. The Hilbert transform was used to extract the instantaneous phase of each signal. The h2, PLV and r2 values range from 0 (independent signals) to 1 (fully correlated signals). C. Data Simulations In order to quantitatively assess the performance of source connectivity approaches, we generated simulated EEG signals following the procedure described in (Cosandier-Rimélé et al. 2008), see figure 1A. The distributed source space consisted in a mesh of the cortical surface (8000 vertices, ~5 mm inter-vertex spacing) that was obtained by segmenting the grey-white matter interface from a normal subject's structural T1-weighted 3D-MRI with Freesurfer (Fischl 2012). Dipoles were located at each vertex of this mesh and oriented radially to the surface at the midway between the white/grey matter interface and the 7 2var(()/())max1var(())xyhytxtytlogijijijxyxyxyxypMIpppijxypijxypijxyppxyMI()()xyittxyPLVe()xt()yt. pial surface. The time-course of each dipole of the source space was generated from a modified version of the physiologically relevant neural mass model reported in (Bourien et al. 2005; Wendling et al. 2002; Wendling et al. 2000). In brief, this computational model was designed to represent a neuronal population with three subsets of neurons (pyramidal cells P and interneurons I and I') interacting via synaptic transmission (Figure 1-A). Pyramidal cells (P) receive endogenous excitatory drive (AMPAergic collateral excitation) from other local pyramidal cells and exogenous excitatory drive from distant pyramidal cells (via external noise input p(t)). They also receive inhibitory drive (GABAergic feedback inhibition) from both subsets of local interneurons (I and I'). In turn, interneurons receive excitatory input (AMPA) from pyramidal cells. A Gaussian noise was used as external input to neuronal population. The mean (m=90) and standard deviation (sigma=30) were adjusted to represent randomly varying density of incoming action potentials (Aps). However, for the purpose of this study, a modification was made to this noise model. Indeed, abrupt increase/decrease of the density of Aps can occur in the external input noise at user-defined times to mimic transient AP volleys from other brain regions involved in the generation of interictal events. Thus, in this model, a simulated IES can be viewed as the responses of a nonlinear dynamical system (comprising positive and negative feedback loops) to transient pulses superimposed on a Gaussian noise (classically used in the neural mass modeling approachs). As in the standard implementation, the shape (spike component followed by a wave component) can still be controlled by adjusting excitation and inhibition parameters of each population (gains in feedback loops). However, the aforementioned modification offers one major advantage: as pulses in the noise input are user-defined, the occurrence times of simulated IESs are controlled, in contrast with the standard implementation where IESs simply result from random fluctuations of the noise. The consequence is that this new model feature allows for simulation multi-focal IESs with well-controlled time shifts. Indeed, as illustrated in Figure 1-A, we could generate delayed epileptiform activity in multiple distant patches just by introducing short delays between the pulses superimposed on the respective input noises of neuronal populations at each patch. Finally, from appropriate setting of the input noise, as well as excitation and inhibition-related parameters at each neural mass included in simulated epileptic sources, a set of epileptiform temporal dynamics was obtained. These dynamics were assigned to a source made of contiguous vertices (active source) manually outlined with a mesh visualization software (Paraview, Kitware Inc., NY, US). Uncorrelated background activities were attributed to the other vertices. Once the amplitude of each elementary dipole was known, EEG simulations were obtained by solving the forward problem in a 3-layer realistic head model (inner skull, outer skull and the scalp with conductivity values of 0.33, 0.0042, 0.33 S/m respectively) using the Boundary Element Method (BEM) with the OpenMEEG (Gramfort et al. 2010) implemented in Brainstorm software. 8 We considered two different scenarios. In the first one (single network), EEG simulations were generated from a single source located in the inferior parietal region (~1000 mm²). In the second one (two interconnected networks) an additional source (~1000 mm²) was placed in the middle temporal gyrus. In that case, the temporal dynamics of the second source were highly correlated with those of the first source, but with a minor delay (30ms). This delay of 30 ms was in the range of 10-50 ms delays that are often observed during interictal spikes at different intracranial recording location (Alarcon et al. 1994; Alarcon et al. 1997; Emerson et al. 1995; Merlet and Gotman 1999) or at different surface sensors (Barth et al. 1984; Ebersole 1994) or between the peaks of dipole source activity (Baumgartner et al. 1995; Merlet and Gotman 1999). This delay was usually interpreted as reflecting propagation between distant regions in the brain. For each scenario, 20 epochs of 60s at 512 Hz containing 30 epileptic spikes were simulated. Each epoch was obtained for a new realization of the input random noise leading to a new realization of epileptic spikes occurring in background activity. Simulated data were imported in Brainstorm for further analysis. Real data Real data were selected from a patient who underwent presurgical evaluation for drug-resistant focal epilepsy. Seizures were stereotyped, with a sudden start, febrile motor automatisms of the upper limb, stretching of the neck and torso and no post- ictal motor deficit. The patient had a comprehensive evaluation including detailed history and neurological examination, neuropsychological testing, structural MRI, standard 32-channels (Micromed, Italy) as well as Dense-EEG 256-channels (EGI, Electrical Geodesic Inc., Eugene, USA) scalp EEG with video recordings and intracerebral EEG recordings (SEEG). MRI showed a focal cortical dysplasia in the mesial aspect of the orbito-frontal region. Dense-EEG was recorded for one hour, at 1000 Hz following the procedure approved by the National Ethics Committee for the Protection of Persons (CPP, agreement number 2012-A01227-36). The patient gave his written informed consent to participate in this study. This recording revealed sub-continuous spike activity at the most left frontopolar basal electrodes. From this interictal epileptic activity, 85 spikes were visually selected away from the occurrence of any artefacts (muscle activity, blood pulsation, eye blinks). Each spike was centered in a 2s window and all 85 windows were concatenated for further analysis. As part of his presurgical evaluation, the patient also underwent intracerebral SEEG recordings with 9 implanted electrodes (10±18 contacts; length: 2 mm, diameter: 0.8 mm; 1.5 mm apart) placed intracranially according to Talairach's stereotactic method in the left frontal and temporal region, see Figure 1C. The positioning of the electrodes was determined from available non-invasive information and hypotheses about the localization of his epileptic zone. From these data, subsets of 25 out of the 118 original leads were selected. This selection was made according to the following criteria: i) only contacts showing grey matter activity were retained and ii) among them, only the contact showing the maximal activity was kept when similar intracerebral activity was observed on several contacts. 9 D. Data analysis Scalp-EEG based interictal epileptic networks. As illustrated in Figure 1B, source activity was estimated using four inverse algorithms (dSPM, wMNE, sLORETA and cMEM, see materials and methods section –A). A baseline of 1s length was used to estimate the noise covariance matrix both on simulated and real scalp EEG data. For real data, source localization was applied on averaged spikes, taking as time reference the maximum of the negative peak, while for simulated data the source localization was applied to non-averaged spikes. The cortical surface was anatomically parcellated into 148 regions of interest (ROI) (Destrieux et al. 2010) and then re-subdivided into ~1500 sub-ROIs using Brainstorm. The 148 ROIs provided initially by the Destrieux Atlas (using Freesurfer) were quasi equally subdivided to obtain the 1500 sub-ROIs with 1cm2 average sizes. Time series of the reconstructed source activities were averaged over each of the 1500 ROIs. After the reconstruction of sources (source localization and estimation of temporal dynamics), functional connectivity was estimated using four methods (r2, h2, PLV, and MI, see materials and methods section -A-). Each quantity was computed on the set of 2 s single spikes. All connectivity matrices (1500 x 1500) were thresholded as follows. We computed the strength of each node of the weighted undirected graph and we kept nodes with the highest 1% strength values. The same threshold was applied on the adjacency matrices for all combinations (inverse/connectivity). The strength was defined as the sum of all edge weights for each node; all weights were positive and normalized between 0 and 1. In order to define the reference networks, all the dipoles were supposed synchronized and the reference network reflected a fully connected undirected graph. In the case of double network scenario, a number of 37 sub-regions (nodes) were considered. The dynamics of the dipoles associated to these nodes were similar and resulting a 37x37 fully connected network where connections (local and remote) between the 37 nodes have the same weight value. Quantification of network similarity. In order to compare the reference brain network simulated in the model with the network identified from simulated scalp EEG by each of the inverse/connectivity combination (Figure 1B), we used a network similarity algorithm recently developed in our team (Mheich et al. 2015a), see supplementary materials for more details about the algorithm. The main advantage of this algorithm is that it takes into account the spatial location (3D coordinates) of the nodes when comparing two networks, in contrast with other methods based on the sole statistical properties of compared graphs. The algorithm provides a normalized Similarity Index (SI): 0 for no similarity and 1 for two identical networks (same properties and topology). The connectivity analysis, the network measures and network visualization were performed using EEGNET (Hassan et al. 2015a; Hassan et al. 2015b). 10 Depth-EEG based interictal epileptic networks. Functional connectivity using h2 were directly computed from SEEG signals at the 25 selected intracerebral electrode contacts. Adjacency matrices (25 x 25) were obtained and thresholded using the same procedure than that applied to the graphs obtained for scalp dense EEG (both simulated and real). Scalp-EEG-based vs. depth-EEG-based epileptic network matching. In order to compare the graphs in the three-dimensional coordinates system of the cortex mesh, the 3D coordinates of the SEEG were first estimated by the co-registering the patient' CT scan and MRI. These points were then projected on the surface mesh. The transformation from MRIs (voxels) coordinates to surface (SCS/MNI) coordinates was realized in brainstorm. The Scalp-EEG-based and depth-EEG-based epileptic networks were visually compared by matching the identified regions (nodes) in both networks. Statistical analysis: On the simulated data, a Wilcoxon rank-sum test was used to compare between the SIs obtained for each combination at each trial, corrected for multiple comparison using Bonferroni approach. Results A. Simulated Data: Influence of the source reconstruction/functional connectivity combination The results obtained in the case of the single network scenario are illustrated in Figure 2, for the 16 different combinations of the source reconstruction and functional connectivity methods. The visual investigation of these results revealed that networks identified using the different combinations of methods were concordant with the reference network (figure 2B). Indeed none of the identified networks had nodes in a remote region (Figure 2A). The qualitative analysis also showed that the number of nodes and the connections between them varied according to the combination of methods used. For a given connectivity approach, changing the localization method did not dramatically modify the network aspect, except for cMEM. Conversely, for a given source localization approach, changing the functional connectivity measure changed, qualitatively, the network. Although this was difficult to assess visually, h2 combined with MNE or sLORETA was giving the network that best matched the reference network while cMEM/MI provided a result that was different from the reference network. Quantification of these differences is provided in figure 2C. Overall, values of network similarity were relatively high and ranged from 70 to 82%. For a given connectivity approach, changing only the localization algorithm slightly modified SI values by 3% (h2) to 8% (MI). For a given source localization approach, the SIs varied within 9% (wMNE) to 12% (dSPM). Results obtained using MI were on average better than PLV, r2 and h2. The combination providing the highest similarity values between the estimated and the actual network was dSPM/MI (82.2%) followed by wMNE/MI (82%) and wMNE-PLV (82%). The lowest similarity value was obtained with the dSPM/h2 combination. The Wilcoxon rank-sum test did not reveal any significant difference between the similarity values obtained in this first study. 11 Results obtained in the case of two interconnected networks for the 16 combinations of the inverse/connectivity methods are reported in Figure 3. Results indicate that the networks identified by all the combinations are relatively close to the model network (figure 3B) since, similarly to the previously scenario, there was no node in other distant regions or in the right hemisphere. The networks did not qualitatively change much for a given connectivity measure except for cMEM. Rather, as observed in the first scenario, the variability between the different combinations was more related to the choice of the connectivity measure, given a source localization approach. The results of PLV (whatever the inverse solution algorithm) provide the closest result to the reference network. cMEM/MI shows also a relatively close network to the reference network while cMEM/h2 indicated, visually, the farthest result from the reference network. Values of network similarity are reported in figure 3C. These values were lower than those in the single network scenario, ranging from 57 to 73%. For a particular connectivity measure, changing the inverse algorithm modified the SIs by 1% (r2) to 8% (h2) while for a given source reconstruction algorithm, the SIs varied between 6% (dSPM) to 13% (wMNE). The combination providing the result closest to the reference network was wMNE/PLV (73%). High values were also obtained with sLORETA/PLV (68%) and cMEM/PLV (66%). The cMEM/ h2 combination shows the lowest SI value (57%). Interestingly, for scenario 2 results obtained with wMNE/PLV were significantly closest to the actual network than the other ones (Wilcoxon rank-sum test, p<0.01, corrected using Bonferroni). B. EEG source localization vs. functional connectivity An essential issue that is addressed in this paper relates to the difference between the proposed "network-based" approach and the classical approach using source localization only. In figure 4, we show two typical examples of the difference between the proposed network-based analysis and the classical localization approach. The first example is for cMEM combined with MI vs. cMEM only. This Figure shows that the information extracted in both cases was noticeably different. The source connectivity approach identified a network close to the reference one (figure 4A), with nodes both in the parietal and in the temporal region (figure 4B). There were no spurious nodes in remote regions. In contrast, with the sole source localization, after averaging the results over a 50 ms interval around each of the epileptic peaks, the parietal source was well retrieved while the temporal source remained almost unobserved. The second example was wMNE/PLV vs. wMNE, the figure shows that the network-based approach was able to identify a network close to the reference with no spurious connections in distant regions. The source localization approach identified the two regions different energies at. Moreover, many spurious sources were observed in remote regions. Similar results were observed for single network configuration. C. Real data: Scalp-EEG-based vs. depth-EEG-based epileptic network 12 The results obtained from real data recorded in a patient are described on Figure 5. In this patient, the sources of scalp EEG interictal spikes were widespread over the left frontal and temporal regions. Sources with maximum activity were found in the left frontal pole and orbitofrontal regions but a substantial activation was also detected in the left temporal as well as right frontal poles (figure 5A, left). When combining wMNE and PLV on the same scalp EEG data, the source connectivity approach retrieved a 5-nodes network in the left frontal lobe, involving the mesial (rectus gyrus) and lateral orbitofrontal region as well as the anterior cingulate gyrus (figure 5A, right). This result was concordant with that the network identified directly from intracerebral recordings by computing the functional connectivity among SEEG signals (Figure 5B right). Indeed, the depth-EEG based network involved six nodes in the left mesial orbito-frontal (rectus gyrus), and anterior cingulate region. All these nodes were also identified by the visual analysis (Figure 5B, left) as regions involved in the main interictal activity (rectus gyrus) as well as in the propagated interictal activity (cingulate gyrus). The similarity indices between networks identified by each of the combination with the depth-EEG-based network are presented in figure 5C. Results showed that the wMNE/PLV provides the highest SI value (70%) followed by wMNE/h2 (47%) and sLORETA/PLV (47%). The cMEM method showed the lowest SI values whatever the connectivity measure (6%, 1%, 1% and 1% for cMEM/MI, cMEM/PLV, cMEM/h2 and cMEM/r2 respectively). Discussion Identifying epileptic brain networks from noninvasive M/EEG data is a challenging issue. Recently, source localization combined with functional connectivity analysis led to encouraging findings in the estimation of functional cortical brain networks from scalp M/EEG recordings (Coito et al. 2015; Jiruska et al. 2013; Malinowska et al. 2014). Nevertheless, the joint use of these two approaches is still novel and raises a number of methodological issues that should be controlled in order to get appropriate and interpretable results. In this paper, we reported a comparative study -in the context of epilepsy- of the networks obtained from all possible combinations between four algorithms to solve the EEG inverse problem and four methods to estimate the functional connectivity. An originality of this study is related to the use of dense EEG signals simulated data from a realistic model of multi-focal epileptic zone as a ground truth for comparing the performance of considered methods. To our knowledge, a model-based evaluation of source connectivity methods has not been performed yet. A second – and still novel - aspect is the use of depth-EEG signals (intracerebral recordings performed during presurgical evaluation of drug-resistant epilepsy) to evaluate the relevance of networks identified from scalp EEG data. Overall results obtained on simulated as well on real data indicated that the combination of the wMNE and the PLV methods leads to the 13 most relevant networks as compared with the ground-truth (simulations) or with the intracerebrally-identified networks (patient data). Results are more specifically discussed hereafter. Methodological considerations The connectivity matrices were thresholded by keeping the nodes with highest strength values (strongest 1%). This procedure was used to standardize the comparison between all the combinations. We were aware about the effect of this threshold and we realized the comparative study using different threshold values. All threshold values were found to lead to the same differences between the method (inverse/connectivity) combinations. In this paper, we have averaged the reconstructed sources within the same region of interest defined by the parcellation process based on Destrieux atlas. This approach was frequently used in the context of M/EEG source connectivity (de Pasquale et al. 2010; Fraschini et al. 2016; Hassan et al. 2015a). However, such an averaging procedure may increase the noise power since its computation is performed over sources that, with some probability, may not exhibit correlation (Brookes et al. 2014) where the need of alternative approaches such as flipping the sign of the sources in each ROIs before averaging the regional time series (Fraschini et al. 2016) or developing methods based on probabilistic maps, a widely approach used in the fMRI-based analysis, for instance. Although EEG source connectivity reduced the problem of volume conduction as compared with scalp EEG connectivity, it does not yet provide a perfect solution. The volume conduction effect is a challenging issue when performing EEG/MEG inverse solution (Schoffelen and Gross 2009). In the connectivity context, the main effect of the volume conduction is the appearance of 'artificial' connections among close sources, a problem often referred to as 'source leakage'. The use of a high spatial resolution (high number of ROIs) may increase this problem. A few approaches have been proposed recently to deal with the source leakage by either normalizing the edges weights by the distance between the nodes or removing the edges between very close sources. Although, these approaches have some advantages, it was shown that, in most cases, they also remove 'real' connections (Schoffelen and Gross 2009). In this context, some connectivity methods such as PLV have been shown to reduce the volume conduction (Hipp et al. 2011). This may explain the good performance of this method. Indeed, in the double network scenario, PLV was able to detect the long-range connections between parietal and temporal networks. Four inverse/connectivity algorithms were evaluated in this paper. It is worth mentioning that some other inverse algorithms like MUSIC-based and beamforming as well some connectivity measures such as power envelope correlation (O'Neill et al. 2015) were not included in this study. Moreover, we were focusing in this paper on evaluating different families of 'functional' connectivity methods regardless the directionality of these connections. Nevertheless, we consider that the analyses of the 'effective' connectivity methods that investigate the causality between brain regions may be of interest in the 14 context of epilepsy (Coito et al. 2016; Coito et al. 2015). In this perspective, methods such as the granger causality, the transfer entropy could be added to expand this comparative study. In addition, all methods evaluated in the paper were bivariate, multivariate methods such as those based on the MVAR model were not included in our study. Different methodological questions appear when using MVAR-based approaches. First, the successful estimation of MVAR such as Partial Directed Coherence (PDC) or Directed Transfer Function (DTF) depends largely on the fitted MAR model, since all information is resulting from the estimated model parameters. In practice, this issue is difficult and directly related to the choice of an optimal model order and an optimal epoch length. Concerning the optimal model, most of the criteria were originally proposed for univariate AR modeling and no consensus was reported about multivariate ones. The second crucial question is how to choose the proper window size specially that MVAR model assumes that the underlying process is stationary, while neurophysiological activity are transient and may rapidly change their states representing high nonstationary behaviors (Pereda et al. 2005). Nevertheless the MVAR (when carefully applied) could provide complementary information not only about the link exists between two signals but also if one structure drive another of if there is feedback between these structures (Kuś et al. 2004). The directionality could be also defined as 'time-delay' between two regional time series which can be computed using linear or nonlinear correlation coefficients. As our main objective in this study was to compare inverse algorithms and 'functional' connectivity methods using same criteria (here we used similarity between reference and estimated undirected graphs), we didn't investigate the time-delays in the presented quantitative analysis. In addition this feature cannot be computed for all the selected methods (the case of the phase synchronization method for instance). We believe that the directionality, estimated from Granger causality or/and time delays, is indeed an interesting supplementary feature in the context of epileptic seizure propagation and will be a potential element for further analysis. The head model used in this study was computed using the Boundary Element Method (BEM) with three layers (skin, skull and brain). This model was widely used in the context of M/EEG source estimation (Fuchs et al. 2007) as a compromise between computational cost and accuracy. Nevertheless, other methods exist to compute the head model such as the Finite Element Method (FEM) or adding other layers such as cerebrospinal fluid (CSF). These methods can possibly have effect of the resultant network (Cho et al. 2015). The evaluation of the above mentioned parameters/factors may be the topic of further investigation. Identification of interictal epileptic networks from scalp dense-EEG data A salient feature of epilepsy in general and epileptic networks is the increased synchronization among interconnected neuronal populations distributed over distant areas. This "hyper"- synchronization often leads to an increase of brain connectivity, not only during the transition to seizures but also during interictal periods, as shown in many studies based on 15 intracranial recordings (see (Wendling et al. 2010) for review). In this context, the combination of the M/EEG source imaging with the functional connectivity measures has recently disclosed promising findings to identify pathological brain networks, at the cortical level (Dai et al. 2012; Lu et al. 2012; Malinowska et al. 2014; Song et al. 2013). However, two factors seem to be crucial for reliable estimation of EEG source connectivity: i) the number of scalp electrodes and ii) the combination between the inverse solution algorithm and the functional connectivity measure. Concerning the number of electrodes, it was reported that the increase of the spatial resolution by using dense EEG may dramatically improve the accuracy of the source localization analysis (Michel and Murray 2012; Song et al. 2015). In addition, the use of dense EEG, as compared to classical montages (32 or 64 electrodes), is needed to accurately identify functional brain networks from scalp EEG (Hassan et al. 2014). To overcome this issue, dense-EEG (256 electrodes) recordings were used in this study. The main feature of this system is the excellent coverage of the subject's head by surface electrodes allowing for improved reconstruction of the cortical activity from non-invasive scalp measurements, as compared with more standard 32- 128 electrode systems (Song et al. 2015). Regarding the combination of inverse/connectivity methods, most of reported studies have empirically selected a combination while this selection was shown to have a dramatic impacts on results, in term of identified network topology (Hassan et al. 2014). The present analysis brings further confirmation of this high variability observed when different inverse solutions and/or connectivity measures are being used in the pipeline leading to cortical networks from EEG signals. A major advantage of the EEG source connectivity approach as presented here is that reconstructed sources and associated networks were identified for the whole brain. Then graph-based metrics (strength values) were computed to characterize the networks and the similarity index was used to compare the networks obtained from various method combinations. In addition, functional connectivity was applied directly to the reconstructed signals and not on derived components. In this regard, this study differs from (Malinowska et al. 2014) where connectivity was estimated on signals components obtained by ICA decomposition. Although the methodological issue of measuring connectivity between independent components still holds, a future interesting study will compare the results obtained from the ICA-based approach to those reported here from source connectivity. EEG source localization vs. functional connectivity Source localization methods have been widely applied to interictal epileptic spikes (Becker et al. 2014). The goal of these approaches is the localization of brain generators of epileptic activity from scalp recordings. A fundamental question that is addressed in this paper relates to the difference between the source connectivity and the source localization approach. This study indicated that the information extracted from dense-EEG recordings in both cases can differ dramatically. First, the 16 connectivity is an additional step to the simple source reconstruction/localization. Second, the fundamental difference between both methods is that the source localization ignores all possible communications between brain regions. When performing source localization analysis, the sources with highest amplitude (averaged at given time period or computed at the instant of peak amplitude of the signal) are classically kept. However, to some extent (depending on threshold), this approach may neglect the possible contribution of "low energy" sources participating into the network activity. Conversely, the hypothesis behind the network-based approach (typically when phase synchronization methods are used as connectivity measure) is that sources can be synchronized regardless their amplitude. To this extent, we believe that the network-based approach allows for revealing networks that are more specific to epileptic networks, as hyper-synchronization phenomena constitute a typical hallmark of such networks. An illustrative example in this paper is the poor involvements of the temporal lobe region when the sole source localization approach (in the case of cMEM) was applied while both parietal and temporal networks (as a priori introduced in the EEG generation model) are retrieved by the connectivity-based approach (cMEM/h2). Note that we have averaged the source localization results in a time window of 50ms to cover the time delay of 30ms set in the model between the two brain regions. Different time window were used to avoid the effect of the window length. All tested windows (30ms, 40ms, 50ms and 60ms) showed similar observations i.e. the absence of the temporal sources (not shown here). The fact that epilepsy is considered as a network disease can explain the low performance of some of the inverse methods as these methods were originally developed to localize 'local' epileptic foci characterized by high- energy sources regardless the interrelationships between brain regions. Our results show that EEG source connectivity methods are more suited in the case of multi-focal epileptic zone. More generally, they support the recent tendency in brain disorder research which is the necessity to move from localizing 'pathological areas' to identifying 'altered networks' (Diessen et al. 2013; Fornito et al. 2015). Epilepsy as a network disorder There is increasing evidence that epileptic activity involves brain networks rather than a single well circumscribed region and that these dysfunctional networks contribute to both ictal and interictal activity (Bourien et al. 2005; Bourien et al. 2004; Coito et al. 2015; Engel Jr et al. 2013; Hipp et al. 2011). Functional connectivity was widely applied to depth-EEG data to predict seizures (Mormann et al. 2000) and identify epileptic networks in partial epilepsies (Bartolomei et al. 2001). These studies showed alterations of synchronization in brain networks during interictal to ictal transition (Wendling et al. 2003) as well as during seizures (Diessen et al. 2013; Jiruska et al. 2013; Schindler et al. 2008). Most of these studies were performed using invasively-recorded data in patient's candidate to surgery. Interestingly, our results show that pathological networks involved during epileptiform activity can also be identified from scalp EEG. 17 Indeed, we have evaluated the performance of a relatively new approach aiming at identifying epileptic brain networks from scalp EEG. The application of the method on real data showed the good performance of this method in term of network identified from scalp EEG as compared with those identified from intracerebral EEG. Note that the comparison was done only by computing h2 between the intracerebral signals based on a large number of studies showing that h2 is one of the most adapted metrics to compute functional connectivity between intracerebral recordings (Bettus et al. 2008; Wendling et al. 2010). Although it is obviously difficult to conclude on a single patient analysis, results showed good matching between scalp-EEG based networks and both the depth-EEG based networks and the expert judgment. Therefore, future work will consist in the application of the EEG source connectivity on a big database of real dense EEG data recorded from epileptic patients. In these patients candidate to surgery, we plan to use also intracerebral EEG signals as a reference to validate the identified networks. In addition, due to the excellent temporal resolution of the EEG, the dynamic behaviors of the epileptic networks will be also explored (Hassan et al. 2015a; Mheich et al. 2015b). Finally, the capacity to describe epileptic activity not only according to the sites where epileptiform activity is generated but also according to the abnormal functional relationships between these sites can definitively improve the surgical approach. We speculate that in order to better understand and ultimately prevent seizures, it is essential to identify and then remove/disconnect pathological nodes of the network (exhibiting abnormal hyper-synchronization capability). The proposed method contributes to this aim and reported results constitute a first step toward the development of more efficient non- invasive diagnostic methods for clinical epileptology. Acknowledgements This work has received a French government support granted to the CominLabs excellence laboratory and managed by the National Research Agency in the "Investing for the Future" program under reference ANR-10-LABX-07-01. It was also financed by the Rennes University Hospital (COREC Project named conneXion, 2012-14). References Adebimpe A, Aarabi A, Bourel-Ponchel E, Mahmoudzadeh M, Wallois F (2016) EEG resting state functional connectivity analysis in children with benign epilepsy with centrotemporal spikes Frontiers in Neuroscience 10 Alarcon G, Guy C, Binnie C, Walker S, Elwes R, Polkey C (1994) Intracerebral propagation of interictal activity in partial epilepsy: implications for source localisation Journal of Neurology, Neurosurgery & Psychiatry 57:435-449 Alarcon G et al. (1997) Origin and propagation of interictal discharges in the acute electrocorticogram. Implications for pathophysiology and surgical treatment of temporal lobe epilepsy Brain 120:2259-2282 Astolfi L et al. (2007) Comparison of different cortical connectivity estimators for high‐resolution EEG recordings Human brain mapping 28:143-157 Babiloni F et al. (2005) Estimation of the cortical functional connectivity with the multimodal integration of high-resolution EEG and fMRI data by directed transfer function Neuroimage 24:118-131 Barth DS, Sutherling W, Engle J, Beatty J (1984) Neuromagnetic evidence of spatially distributed sources underlying epileptiform spikes in the human brain Science 223:293-296 18 Bartolomei F, Wendling F, Bellanger J-J, Régis J, Chauvel P (2001) Neural networks involving the medial temporal structures in temporal lobe epilepsy Clinical neurophysiology 112:1746-1760 Baumgartner C et al. (1995) Propagation of interictal epileptic activity in temporal lobe epilepsy Neurology 45:118-122 Becker H et al. (2014) EEG extended source localization: tensor-based vs. conventional methods NeuroImage 96:143-157 Betti V, Della Penna S, de Pasquale F, Mantini D, Marzetti L, Romani GL, Corbetta M (2013) Natural scenes viewing alters the dynamics of functional connectivity in the human brain Neuron 79:782-797 Bettus G, Wendling F, Guye M, Valton L, Regis J, Chauvel P, Bartolomei F (2008) Enhanced EEG functional connectivity in mesial temporal lobe epilepsy Epilepsy Res 81:58-68 doi:10.1016/j.eplepsyres.2008.04.020 Bola M, Sabel BA (2015) Dynamic reorganization of brain functional networks during cognition NeuroImage 114:398-413 Bourien J, Bartolomei F, Bellanger JJ, Gavaret M, Chauvel P, Wendling F (2005) A method to identify reproducible subsets of co-activated structures during interictal spikes. Application to intracerebral EEG in temporal lobe epilepsy Clinical neurophysiology : official journal of the International Federation of Clinical Neurophysiology 116:443-455 doi:10.1016/j.clinph.2004.08.010 Bourien J, Bellanger JJ, Bartolomei F, Chauvel P, Wendling F (2004) Mining reproducible activation patterns in epileptic intracerebral EEG signals: application to interictal activity IEEE transactions on bio-medical engineering 51:304-315 doi:10.1109/TBME.2003.820397 Brookes MJ et al. (2014) Measuring temporal, spectral and spatial changes in electrophysiological brain network connectivity Neuroimage 91:282-299 Caparos M, Louis-Dorr V, Wendling F, Maillard L, Wolf D (2006) Automatic lateralization of temporal lobe epilepsy based on scalp EEG Clinical neurophysiology 117:2414-2423 Cho J-H, Vorwerk J, Wolters CH, Knösche TR (2015) Influence of the head model on EEG and MEG source connectivity analysis Neuroimage 110:60-77 Chowdhury RA, Lina JM, Kobayashi E, Grova C (2013) MEG source localization of spatially extended generators of epileptic activity: comparing entropic and hierarchical Bayesian approaches PloS one 8:e55969 Clarke C, Janday B (1989) The solution of the biomagnetic inverse problem by maximum statistical entropy Inverse Problems 5:483 Coito A et al. (2016) Altered directed functional connectivity in temporal lobe epilepsy in the absence of interictal spikes: A high density EEG study Epilepsia Coito A et al. (2015) Dynamic directed interictal connectivity in left and right temporal lobe epilepsy Epilepsia 56:207-217 Cosandier-Rimélé D, Merlet I, Badier J-M, Chauvel P, Wendling F (2008) The neuronal sources of EEG: modeling of simultaneous scalp and intracerebral recordings in epilepsy NeuroImage 42:135-146 Dai Y, Zhang W, Dickens DL, He B (2012) Source connectivity analysis from MEG and its application to epilepsy source localization Brain topography 25:157-166 Dale AM, Liu AK, Fischl BR, Buckner RL, Belliveau JW, Lewine JD, Halgren E (2000) Dynamic statistical parametric mapping: combining fMRI and MEG for high-resolution imaging of cortical activity Neuron 26:55-67 David O, Cosmelli D, Hasboun D, Garnero L (2003) A multitrial analysis for revealing significant corticocortical networks in magnetoencephalography and electroencephalography Neuroimage 20:186-201 David O, Garnero L, Cosmelli D, Varela FJ (2002) Estimation of neural dynamics from MEG/EEG cortical current density maps: application to the reconstruction of large-scale cortical synchrony IEEE Transactions on Biomedical Engineering, 49:975-987 de Pasquale F et al. (2010) Temporal dynamics of spontaneous MEG activity in brain networks Proceedings of the National Academy of Sciences 107:6040- 6045 Destrieux C, Fischl B, Dale A, Halgren E (2010) Automatic parcellation of human cortical gyri and sulci using standard anatomical nomenclature Neuroimage 53:1-15 Diessen E, Diederen SJ, Braun KP, Jansen FE, Stam CJ (2013) Functional and structural brain networks in epilepsy: what have we learned? Epilepsia 54:1855-1865 Ding L, Worrell GA, Lagerlund TD, He B (2007) Ictal source analysis: localization and imaging of causal interactions in humans Neuroimage 34:575-586 Ebersole J (1994) Non‐invasive localization of the epileptogenic focus by EEG dipole modeling Acta Neurologica Scandinavica 89:20-28 Emerson RG, Turner CA, Pedley TA, Walczak TS, Forgione M (1995) Propagation patterns of temporal spikes Electroencephalography and clinical neurophysiology 94:338-348 Engel Jr J, Thompson PM, Stern JM, Staba RJ, Bragin A, Mody I (2013) Connectomics and epilepsy Current opinion in neurology 26:186 Fischl B (2012) FreeSurfer Neuroimage 62:774-781 Fornito A, Zalesky A, Breakspear M (2015) The connectomics of brain disorders Nature Reviews Neuroscience 16:159-172 Fraschini M, Demuru M, Crobe A, Marrosu F, Stam CJ, Hillebrand A (2016) The effect of epoch length on estimated EEG functional connectivity and brain network organisation Journal of Neural Engineering 13:036015 Fuchs M, Wagner M, Kastner J (2007) Development of volume conductor and source models to localize epileptic foci Journal of Clinical Neurophysiology 24:101-119 Gotman J (1983) Measurement of small time differences between EEG channels: method and application to epileptic seizure propagation Electroencephalography and clinical neurophysiology 56:501-514 Gotman J (1987) Interhemispheric interactions in seizures of focal onset: data from human intracranial recordings Electroencephalography and clinical neurophysiology 67:120-133 Gramfort A, Papadopoulo T, Olivi E, Clerc M (2010) OpenMEEG: opensource software for quasistatic bioelectromagnetics Biomed Eng Online 9:45 Hämäläinen M (2005) MNE software user's guide NMR Center, Mass General Hospital, Harvard University 58:59-75 Hämäläinen MS, Ilmoniemi R (1994) Interpreting magnetic fields of the brain: minimum norm estimates Medical & biological engineering & computing 32:35-42 Hassan M, Benquet P, Biraben A, Berrou C, Dufor O, Wendling F (2015a) Dynamic reorganization of functional brain networks during picture naming Cortex 73:276-288 Hassan M, Dufor O, Merlet I, Berrou C, Wendling F (2014) EEG Source Connectivity Analysis: From Dense Array Recordings to Brain Networks PloS one 9:e105041 Hassan M, Shamas M, Khalil M, El Falou W, Wendling F (2015b) EEGNET: An Open Source Tool for Analyzing and Visualizing M/EEG Connectome PloS one 10:e0138297 Hassan M, Wendling F (2015) Tracking dynamics of functional brain networks using dense EEG IRBM 36:324-328 Hipp JF, Engel AK, Siegel M (2011) Oscillatory synchronization in large-scale cortical networks predicts perception Neuron 69:387-396 Hoechstetter K, Bornfleth H, Weckesser D, Ille N, Berg P, Scherg M (2004) BESA source coherence: a new method to study cortical oscillatory coupling Brain topography 16:233-238 Jiruska P, de Curtis M, Jefferys JG, Schevon CA, Schiff SJ, Schindler K (2013) Synchronization and desynchronization in epilepsy: controversies and hypotheses The Journal of physiology 591:787-797 doi:10.1113/jphysiol.2012.239590 Kramer MA, Cash SS (2012) Epilepsy as a disorder of cortical network organization The Neuroscientist 18:360-372 19 Kuś R, Kamiński M, Blinowska KJ (2004) Determination of EEG activity propagation: pair-wise versus multichannel estimate IEEE Transactions on Biomedical Engineering, 51:1501-1510 Laufs H (2012) Functional imaging of seizures and epilepsy: evolution from zones to networks Current opinion in neurology 25:194-200 Le Van Quyen M, Soss J, Navarro V, Robertson R, Chavez M, Baulac M, Martinerie J (2005) Preictal state identification by synchronization changes in long-term intracranial EEG recordings Clinical Neurophysiology 116:559-568 Liljeström M, Kujala J, Stevenson C, Salmelin R (2015) Dynamic reconfiguration of the language network preceding onset of speech in picture naming Human brain mapping 36:1202-1216 Lu Y, Yang L, Worrell GA, He B (2012) Seizure source imaging by means of FINE spatio-temporal dipole localization and directed transfer function in partial epilepsy patients Clinical Neurophysiology 123:1275-1283 Malinowska U, Badier JM, Gavaret M, Bartolomei F, Chauvel P, Bénar CG (2014) Interictal networks in magnetoencephalography Human brain mapping 35:2789-2805 Merlet I, Gotman J (1999) Reliability of dipole models of epileptic spikes Clinical Neurophysiology 110:1013-1028 Mheich A, Hassan M, Gripon V, Dufor O, Khalil M, Berrou C, Wendling F A novel algorithm for measuring graph similarity: application to brain networks. In: 7th International IEEE EMBS Neural engineering Conference, Montpellier, France, 2015a. Mheich A, Hassan M, Khalil M, Berrou C, Wendling F (2015b) A new algorithm for spatiotemporal analysis of brain functional connectivity Journal of neuroscience methods 242: 77–81 Michel CM, Murray MM (2012) Towards the utilization of EEG as a brain imaging tool Neuroimage 61:371-385 Mormann F, Lehnertz K, David P, Elger CE (2000) Mean phase coherence as a measure for phase synchronization and its application to the EEG of epilepsy patients Physica D: Nonlinear Phenomena 144:358-369 O'Neill GC, Barratt EL, Hunt BA, Tewarie PK, Brookes MJ (2015) Measuring electrophysiological connectivity by power envelope correlation: a technical review on MEG methods Physics in medicine and biology 60:R271 Pascual-Marqui RD (2002) Standardized low-resolution brain electromagnetic tomography (sLORETA): technical details Methods Find Exp Clin Pharmacol 24:5-12 Pereda E, Quiroga RQ, Bhattacharya J (2005) Nonlinear multivariate analysis of neurophysiological signals Progress in neurobiology 77:1-37 Ramon C, Holmes MD (2013) Noninvasive localization of epileptic sites from stable phase synchronization patterns on different days derived from short duration interictal scalp dEEG Brain topography 26:1-8 Schindler KA, Bialonski S, Horstmann MT, Elger CE, Lehnertz K (2008) Evolving functional network properties and synchronizability during human epileptic seizures Chaos 18:033119 doi:10.1063/1.2966112 Schoffelen JM, Gross J (2009) Source connectivity analysis with MEG and EEG Human brain mapping 30:1857-1865 Song J et al. (2015) EEG source localization: Sensor density and head surface coverage Journal of neuroscience methods 256:9-21 Song J, Tucker DM, Gilbert T, Hou J, Mattson C, Luu P, Holmes MD (2013) Methods for examining electrophysiological coherence in epileptic networks Frontiers in neurology 4 Tadel F, Baillet S, Mosher JC, Pantazis D, Leahy RM (2011) Brainstorm: a user-friendly application for MEG/EEG analysis Computational intelligence and neuroscience 2011:8 van Mierlo P, Papadopoulou M, Carrette E, Boon P, Vandenberghe S, Vonck K, Marinazzo D (2014) Functional brain connectivity from EEG in epilepsy: Seizure prediction and epileptogenic focus localization Progress in neurobiology 121:19-35 Vecchio F et al. (2014) Cortical connectivity in fronto-temporal focal epilepsy from EEG analysis: A study via graph theory Clinical Neurophysiology Wendling F, Ansari-Asl K, Bartolomei F, Senhadji L (2009) From EEG signals to brain connectivity: a model-based evaluation of interdependence measures Journal of neuroscience methods 183:9-18 Wendling F, Bartolomei F, Bellanger J-J, Bourien J, Chauvel P (2003) Epileptic fast intracerebral EEG activity: evidence for spatial decorrelation at seizure onset Brain 126:1449-1459 Wendling F, Bartolomei F, Bellanger JJ, Chauvel P (2002) Epileptic fast activity can be explained by a model of impaired GABAergic dendritic inhibition The European journal of neuroscience 15:1499-1508 Wendling F, Bellanger J-J, Bartolomei F, Chauvel P (2000) Relevance of nonlinear lumped-parameter models in the analysis of depth-EEG epileptic signals Biological cybernetics 83:367-378 Wendling F, Chauvel P, Biraben A, Bartolomei F (2010) From intracerebral EEG signals to brain connectivity: identification of epileptogenic networks in partial epilepsy Frontiers in systems neuroscience 4 20 21 Figure 1: Structure of the investigation. A) Simulated epileptic spikes: Model used to generate epileptic spikes (see Materials and methods/Data/Simulations section for detailed description), B) Identification of interictal epileptic network: First, a network is generated by the model and considered as the 'ground truth'. By solving the forward problem, synthetic dense EEG data are generated. These signals are then used to solve the inverse problem in order to reconstruct the dynamics of sources using three different inverse solutions (wMNE, sLORETA, dSPM and cMEM). The statistical couplings are then computed between the reconstructed sources using three different methods (r2, PLV, h2 and MI). The identified network by each combination (inverse/connectivity) was then compared with the original network using a 'network similarity' algorithm [13] and C) Intracerebral recordings: The positions of the intracerebral SEEG signals used in the real application. The corrdinates of the electrode's contacts was obtianed by the CT/MRI coregistration. Abbreviations: wMNE: weighted Minimum Norm Estimate; sLORETA: Standardized low resolution brain electromagnetic tomography; dSPM: dynamical Statistical Parametric Mapping; cMEM: standard Maximum Entropy on the Mean; r2: linear correlation coefficient; PLV: Phase Locking Value; h2: nonlinear correlation coefficient; MI: mutual information, P: pyramidal cells, I: Inhibitory interneurons. 22 23 Figure 2: One network scenario. A) Brain networks obtained by using the different inverse and connectivity methods, B) The original network (ground truth) is shown and C) Values (mean ± standard deviation) of the similarity indices computed between the network identified by each combination and the model network. 24 25 Figure 3: Two networks scenario. A) Brain networks obtained by using the different inverse and connectivity methods. B) The original network (ground truth) is shown and C) Values (mean ± standard deviation) of the similarity indices computed between the network identified by each combination and the model network. 26 27 Figure 4: Source localization vs. source connectivity. A. the reference network. B. Results obtained by the network-based approach (cMEM/MI and wMNE/PLV) and C. Results obtained by the localization based approach (cMEM and wMNE) using same window of analysis. Results were averaged over a 50 ms interval around each of the spike peaks. Red color represents the sources with the highest energy. 28 29 Figure 5: Application on real data. A) Scalp EEG: Results of the source localization approach using wMNE (right) and source connectivity using wMNE/PLV (left), B. Intracerebral EEG: Regions visually identified by the epileptologist (right) and the network obtained by computing the functional connectivity between the intracerebral EEG signals (left), C. Similarity indices: The SI values obtained between the network identified by each of the combination and the intracerebral- EEG-based network. 30
1702.08421
1
1702
2017-02-27T18:27:48
The role of the observer in goal-directed behavior
[ "q-bio.NC" ]
In goal-directed behavior, a large number of possible initial states end up in the pursued goal. The accompanying information loss implies that goal-oriented behavior is in one-to-one correspondence with an open subsystem whose entropy decreases in time. Yet ultimately, the laws of physics are reversible, so systems capable of yielding goal-directed behavior must transfer the information about initial conditions to other degrees of freedom outside the boundaries of the agent. To operate steadily, they must consume ordered degrees of freedom provided as input, and be dispensed of disordered outputs that act as wastes from the point of view of the aimed objective. Here I argue that a physical system may or may not display goal-directed behavior depending on what exactly is defined as the agent. The borders of the agent must be carefully tailored so as to entail the appropriate information balance sheet. In this game, observers play the role of tailors: They design agents by setting the limits of the system of interest. Brain-guided subjects perform this creative observation task naturally, implying that the observation of goal-oriented behavior is a goal-oriented behavior in itself. Minds evolved to cut out pieces of reality and endow them with intentionality, because ascribing intentionality is an efficient way of modeling the world, and making predictions. One most remarkable agent of whom we have indisputable evidence of its goal-pursuing attitude is the self. Notably, this agent is simultaneously the subject and the object of observation.
q-bio.NC
q-bio
The role of the observer in goal-directed behavior In´es Samengo 7 1 0 2 b e F 7 2 ] . C N o i b - q [ 1 v 1 2 4 8 0 . 2 0 7 1 : v i X r a Abstract In goal-directed behavior, a large number of possible initial states end up in the pursued goal. The accompanying information loss implies that goal-oriented behavior is in one-to-one corre- spondence with an open subsystem whose entropy decreases in time. Yet ultimately, the laws of physics are reversible, so entropy variations are necessarily a consequence of the way a system is described. In order to reconcile different levels of description, systems capable of yielding goal-directed behavior must transfer the information about initial conditions to other degrees of freedom outside the boundaries of the agent. To operate steadily, they must consume ordered degrees of freedom provided as input, and be dispensed of disordered outputs that act as wastes from the point of view of the aimed objective. Broadly speaking, hence, goal-oriented behav- ior requires metabolism, even if conducted by non-living agents. Here I argue that a physical system may or may not display goal-directed behavior depending on what exactly is defined as the agent. The borders of the agent must be carefully tailored so as to entail the appropriate information balance sheet. In this game, observers play the role of tailors: They design agents by setting the limits of the system of interest. Their computation may be iterated to produce a hierarchy of ever more complex agents, aiming at increasingly sophisticated goals, as observed in darwinian evolution. Brain-guided subjects perform this creative observation task naturally, implying that the observation of goal-oriented behavior is a goal-oriented behavior in itself. Minds evolved to cut out pieces of reality and endow them with intentionality, because ascrib- ing intentionality is an efficient way of modeling the world, and making predictions. One most remarkable agent of whom we have indisputable evidence of its goal-pursuing attitude is the self. Notably, this agent is simultaneously the subject and the object of observation. 1 A bunch of nucleic acids swim among many other organic compounds forming a cytoplas- matic soup, and somehow, manage to arrange themselves into precisely the sequence required for DNA replication. Carbon dioxide molecules steadily stick to one another materializing a solid tree trunk out of a tiny seed. Owls eat the young bats with poor navigation ability, thereby improving the eco-location proficiency of the species. The neurons in a dog's brain fire pre- cisely in the required sequence to have the dog bury its bone, hiding it from other dogs. The wheels, breaks, and clutch of a self-driving car coordinate their actions in order to reach the parking area of a soccer field, no matter the initial location of the car, nor the traffic along the way. The limbs of the Argentine soccer players display a complex pattern of movements that carry the ball, through kicks and headers, at Messi's feet in front of the keeper... kick... goal! This essay is about goals. In all the above examples, a collection of basic elements, follow- ing local and apparently purpose-less laws, manage to steer the value of certain variables into some desired regime. The initial state is rather arbitrary, and yet, the agents manage to adap- tively select, out of many possible actions, the maneuvers that are suited to conduct the system to the desired goal. Throughout these seemingly intelligent choices, order appears to raise from disorder. Scattered nucleotides become DNA. Air and dust become trees. Owl hunger becomes sophisticated eco-location organs. Neural activity becomes a buried bone. A car anywhere in the city becomes a car at a specific location. A football anywhere in the stadium becomes a football in the goal. How do the components of each system know what to do, and what not to do, in order to reach the goal? This is the question that will entertain us here. Initial G G A T C T A C C C T A G A G T A T C G C T A G Final G A A T C T G C Figure 1: In DNA replication, many initial states are mapped onto a single final state. The entropy is therefore high at the beginning and low at the end. One important characteristic of goal- directed agents is that they are flexible: They reach the goal from multiple initial condi- tions, and are typically able to circumvent ob- stacles. For example, in DNA replication, the initial state is one out of many configurations in which nucleotides can be spatially dis- tributed in a solution of organic compounds. The final state, the goal, is the precise spatial arrangement of those same nucleotides within the newly constructed DNA strand. In the soccer stadium, the ball may be initially in any location, the final state is the ball at the goal. Multiple initial states are hence mapped onto a single final state, as in Fig. 1. In physical terms, the non-injective nature of this mapping implies a reduction in entropy. 2 Admittedly, the final state need not be strictly unique. In DNA replication, permutations of equal nucleotides are still allowed in the final state, and occasionally, there might also be a few errors in the replication process. Dogs may consider more than a single location for the concealed bounty, and Messi may choose to shoot the ball anywhere inside the 24 ft wide by 8 ft high of the goal. Such restricted amounts of freedom, or even the occasional failures to reach the final state (shooting an own goal, for example), by no means compensate the abrupt reduction in entropy that takes place throughout the process. In fact, were entropy not to decrease, the system would not exhibit goal-oriented behavior. We are used to associating entropy increments with information losses, and entropy re- ductions with information gains. Here I am taking the opposite view: Entropy reductions are associated with information losses. The two views are not incompatible, they simply refer to different things. The first case deals with a closed system and information about macroscopic variables. The second, with an open system and microscopic variables. When a closed macro- scopic system evolves in the direction that maximizes the entropy of all compatible microscopic states (the usual case in closed thermodynamical systems), the final macro-state does not allow us to deduce the initial macro-state, since the mapping between them is non-injective. Were we to know the detailed final micro-state, however, we would be able to deduce the initial micro- state. When a goal-oriented system evolves in the direction of decreasing entropy, the final micro-state does not allow us to deduce the initial micro-state, but for a different reason: goal- oriented systems are open, and they interact with degrees of freedom we are not keeping track of. In this essay, the distinction between micro and macro-states is not emphasized, because the phenomena we deal with are not always divisible into separate scales. The notion of goal-oriented behavior that is used here always brings about an entropy reduc- tion. I now want to demonstrate the reciprocal statement: If a system reduces its entropy, a goal can be ascribed to the process. Therefore, entropy reduction and goal-oriented behavior are in a one-to-one correspondence. The goal in question can always be defined by the restricted set of values that the variables acquire in the final state: the target DNA sequence, the buried bone, the ball at the goal. Of course, the reduction in entropy must first be verified: a broad set of initial states must evolve into a small final set. A car that in a single trial travels from one location to another is not guaranteed to be a self-driving car. Only if the initial location has proven to be arbitrary, and the traffic conditions variable, can goal-directed behavior be arrogated. The notion of entropy is subtle, since it not only characterizes a physical system, but also, the way it is described. When the universe is described at its utmost basic level (assuming there is one such level), all we have is a collection of fundamental particles evolving from some initial state. If the state of all particles is specified, the total entropy of the universe vanishes. Time 3 reversibility of the laws of physics dictates entropy to remain zero for all past and future times. Therefore, there is no way to attain neither an increase nor a reduction in entropy. Energy dissipation and goal-directed behavior, hence, are absent from the complete description. We need to blur our point of view to give them a chance, either by restricting the description to macroscopic variables, or to subsystems. In fact, the main conclusion of this essay is that an observer with a very special point of view is required for agency to exist. If the information about the initial conditions is apparently lost in goal-oriented behavior, then such information must be somehow concealed in degrees of freedom we are not keeping track of. They may have been moved into too microscopic states to be monitored, or into fluid degrees of freedom that, by the time the goal is reached, have already exited the subsystem under study. What we track, and what we ignore, hence, plays a crucial role in agency. hole sliding door Figure 2: A demon controls the sliding door, al- lowing particles to pass from right to left, but not the other way round. The initial state of ev- ery molecule that the demon has already acted on (letting it pass or not) is recorded in its mem- ory, and depicted in blue. To be consistent with the second law of thermodynamics, processes where entropy decreases are only possible in open sys- tems that somehow interact with the external world. Originally, they were supposed to re- quire an energy influx. This is, however, not a necessary condition: Sometimes, the sole exchange of information suffices. A good ex- ample is Maxwell's Demon [1]. Suppose we have a gas enclosed in two adjacent chambers communicated by a small hole in the wall be- tween them (Fig. 2). The hole may or may not be covered by a sliding door controlled by a demon. Initially, both chambers have equal pressure and density. The demon then opens or closes the hole selectively, depend- ing on whether a molecule approaches from one side, or the other. Molecules coming from the right are allowed to pass into the left cham- ber, but not the other way round. As time goes by, molecules accumulate on the left side, eventually leaving the right side empty. The collection of all gas molecules can be interpreted as performing goal-directed behavior: No matter the initial state, gas is gradually compressed into the left chamber. This final state can be conceived as a goal, and it comprises a reduction in entropy: initially each particle can be anywhere in the two chambers, and in the final state, they are all in the left side. Arrogating purpose, in this case, is to assume that the gas -- who takes the role of the agent -- wants to shrink. Other verbs may be used (tends to, is inclined to, 4 etc.), but the phrasing is irrelevant. As uncanny as it may seem, arrogating purpose to the gas is a rather accurate description of the gas' phenomenology. The gas + demon is a toy model of a closed system, so no interaction with the outside world is allowed. To perform the task, the demon needs to acquire information on the location of each molecule approaching the hole, to then decide whether to let it pass or not. In a slightly modified version of this system, Bennett [2] demonstrated that the storage of information in the demon's memory can be done with no energy expenditure, as long as the memory is initially blank, and there is plenty of storage capacity. The work required to move and stop the door, as well as the energy needed to measure the position of particles and to maintain the demon alive, can also be made as small as desired, simply diminishing mechanical friction, and moving slowly. The demon is however not allowed to delete the acquired information, because information erasure requires energy consumption, at a minimal cost of kBT per erased bit [3]. Therefore, as time goes by, the information of the initial location of each gas molecule is erased from the gas, and copied onto the demon's memory. The gas gradually reduces its entropy only if we are careful to exclude the demon from what we define as the system. If we include the demon (and its memory), entropy simply remains constant, since all the details of the initial state are still stored. Depending on the observer's choices, then, entropy may or may not decrease, meaning that arrogating agency may or may not be possible. A subsystem can only decrease its entropy if it somehow gets rid of initial conditions. In DNA replication, after the addition of each new nucleotide to the developing strand, the ini- tial location of the free nucleotide determines the final configuration of the mediating enzimes, thereby transferring the information of the initial state to a change in the 3-dimensional configu- ration of nearby proteins. If enzimes are not restored to their functional state, the process cannot be iterated. So enzimes, in turn, must pass the information on somewhere else. This transfer is actually the important point in the emergence of goal-directed behavior. Energy consumption is only helpful if energy is degraded in the process: ordered energy sources must be transformed into disordered products. In animal cells, order arrives as glucose and oxygen molecules. Dis- order exits as carbon dioxide, water and faster molecular motion (heat). The input degrees of freedom, specifically in the case of glucose, are conformed of atoms tidily organized into large molecules. The output degrees of freedom are transported by smaller molecules, amenable to be arranged in many more configurations. The laws of physics are ultimately reversible, so initial conditions cannot be truly erased, they can only be shuffled around. As an example, Edward Fredkin studied how non-dissipative systems, such as our universe, may perform the usual logical computations (AND, OR, etc), 5 which are themselves non-invertible [4]. We know that 0 AND 1 = 0. However, knowing that the result of the operation is 0 does not suffice to identify the two input variables. If the compu- tation is performed by an ultimately non-dissipative system, the information of the initial input variables must be somehow moved into some other variable, albeit perhaps not in a manner that is easily accessible. Fredkin's solution was to prove that computing required some extra input variables, not needed for the computation per se, but mandatory for the information balance. When performing a single logical operation (say, for example, AND), the additional variables are in a well defined state (no uncertainty), and throughout the computation, they acquire so- called garbage values (garbage because they are not required to perform the computation), that represent those input degrees of freedom that cannot be deduced from the output. Copying part of the input into garbage variables ensures that no information is lost, and the computation becomes feasible in a non-dissipative physical substrate. Input Output Figure 3: Goal-directed systems (blue ball) eat up ordered degrees of freedom, and produce disordered degrees of freedom. Ascribing agency is all about ignoring who really did the job (the Universe, to put it grandly), and arrogating intentionality to an entropy-reducing subsystem. The task of the observer is to design the borders of the sub- system so as to allow ordered degrees of free- dom to be progressively incorporated, and/or disordered ones to be eliminated. If the goal is to be achieved repeatedly, a steady flow of order is required, as well as a regular garbage collection service. Purposeful agents, hence, only emerge from sub systems that eat up order (Fig. 3). Broadly speaking, they can be said to breathe, or to be endowed with metabolism, even if they need not be alive in the biological sense. Maxwell's demon hid the initial conditions of the gas in its memory. The dog, the self- driving car, and the soccer players, all hide their own initial state and that of the environment inside their memories. Memories can of course be erased, but erasures consume energy, and they are ultimately no more than flushing initial conditions into the high-entropy products of energy degradation. For a long time, scientists failed to include memories as part of the sys- tems under study, so goal-directed behavior sometimes appeared paradoxical. Here I argue that observers attribute agency by disregarding initial conditions. Of course, observers are free to delineate the borders of the subsystem under study as they wish. They can always shape the limits of what they define as the agent in such a way as to have it do all sorts of wonderful things, as achieve goals, and reduce entropy. The agent must be fed with order, and the mess must be cleaned up, but still, it can be done. The natural question is therefore: What is in- 6 teresting in goal-directed behavior if the observer is allowed to engineer the very definition of the agent, in order to get the desired result? Plants grow because what we define as a plant is the stuff that grows every spring, and not the dirt left on the ground every autumn. Species improve because we restrict the definition of a species to the material that a posteriori is seen as successful, and exclude the corpses left behind of those who failed. Cell division seems to be a productive business because the waste products are not defined to be part of cells. Returning to the question posed above "How do the components of the system know what to do, and what not to do, in order to reach the goal?", we can now provide an answer. Components know nothing, observers do. Just as photographers select an arbitrary plane in the visual world where to focus an image and engender a sharp object (Fig. 4), so do observers choose which variables compose the system, and which do not, so that a goal emerges. Should we be amazed that the world we live in allows observers to create agents? Could we not live in a universe where as- signing agency were downright impossible? I would be very much surprised if it were so. The impossibility to define goal-directed be- havior would mean that no subsystems exist where entropy decreases. The global entropy growth that takes place in the whole universe should develop uniformly and monotonously all throughout space and time, allowing for no local oscillations. That would be ordered in- deed! I do not expect disorder to arise in such an orderly manner. Figure 4: Observers, just as photographers, se- lectively focus on an aspect of reality, to satisfy their cognitive appetite [6]. Potograph kindly supplied by Luc´ıa Samengo. Within this picture, all the interesting events seem to take place in the observer's creative act. Any local decrement of entropy, no matter how trivial, appears to suffice for an observer to be able to ascribe agency. We are demanding little of the world, and a lot of the observer. But does the evolving world not have organizational merits of its own? If we look at the history of events taking place in our planet, as time goes by, agents seem to become increasingly sophisticated. Compare the strategy of a replicating DNA molecule in a bacterium with the one of Menelaus of Sparta to recover Helen of Troy, and thereby, ultimately manage bisexual reproduction. All the complexity of the bacterium's strategy is present in Menelaus', but not the other way round. Evolution seems to be striving towards what appears to be a runaway escalation of sophistica- tion and design [5]. Is the development of refined agents something that only depends on the observer's creativity, or is it something actually taking place independently of observers? 7 Merits are shared, I believe. Observers produce agents. In the absence of agents, no sub- systems are cut out of the wholeness of the cosmos, and complexity cannot be measured. Once observers are in play, even if they might have never intended it, it turns out that the computa- tion they perform is liable to iteration. Agency allocation implies that many equal final states are produced from many different initial states. The final states are similar to one another, and similarity is a form of order. Such final states can therefore be used as the ordered degrees of freedom that a higher-level process may use as fuel. This fuel needs to enter into a noisy system for higher-level agents to emerge (Fig. 5). Noise is typically instantiated by a changing environment, often in combination with the occasional mistakes that may have happened in the lower-level process, as mutations in DNA replication. By iterating the algorithm, profuse RNA replication in free solutions can be observed to give rise to prokaryote cells, who in turn evolve into eukaryotes, from which multi-cellular organisms appear, all the way up to the ever growing branches of the tree of life. In the way, conscious humans, civilization, and artificial intelligence emerge. As well as a lot of garbage, as environmentalists wisely remind us. Final state of 1st level, and ordered fuel of 2nd level Initial state (1st level) l e u f d e r e d r O ) l e v e l t s 1 ( Initial state (2nd level) Final state of 2nd level and ... The tree of life develops as a continuous process, irrespectively of whether observers interpret it or not. Observers are required to dissect it into agents, and to evaluate their so- phistication. As we climb the ladder of evo- lutionary design, the fuel degrees of freedom diminish in number, and increase in complex- 1st level wastes 2nd level wastes ity, the latter defined as the number of bits re- Figure 5: The goal achieved by low-level agents can become the ordered fuel of higher-level agents. quired to describe their inner structure. If the resources at the bottom level are finite, the process cannot be iterated indefinitely, since eventually, too few and too complex degrees of freedom may not be identifiable as multiple instances of one single ordered pattern. Brain-guided observers are continuously ascribing agency. They do so because the role of a brain is to model the world around its carrier, so that effective survival strategies can be implemented. They have evolved to do so. Mental models must capture the regularities of the world, and discard the noise. Here, noise is defined as the degrees of freedom that are irrelevant to predicting those features of the environment that affect the observer's fitness. It would be a waste of resources, if not impossible, for us to represent in our brains all what happens in a dog's brain. Much more efficient is to ascribe agency, and conclude that the dog wants to bury the bone. We cannot follow the evolution of all the bats that were eaten by owls, we therefore conclude that the predation of owls sharpens the eco-location capacity of bats. We do 8 not care for the details with which self-driving cars are programmed, we just think of them as goal-directed. We need an economic description, so we assign agency. Observers do not assign agency to all the entropy-reducing systems they meet. Purpose is only arrogated to subsystems for which there is no evident source of order, or for sources that are too costly to represent. The cost of a representation is judged in terms of its contribution to prediction accuracy. If we only look at the gas controlled by Maxwell's demon, ignoring the demon itself, we conclude that the gas wants to shrink. If, however, the demon takes weekends off, the purposeful model of the gas loses accuracy. A more sophisticated representation dis- cerning between week days and weekends is needed. Assigning agency may or may not be a convenient strategy, depending on the trade-off between the economy of the representation and the prediction errors it induces. Arrogating agency in excess, for example by believing that all what happens is maneuvered by some obscure intentionality, yields a poor prediction strategy. Observation is the result of development: Observers learn how to observe, and they do so within the framework of learning theory [7]. They are first exposed to multiple examples of the process, that act as the training set. Before learning, the final state can only be predicted from the initial one if all degrees of freedom are tracked - a representation capacity that observers typically lack. Making the best use of their resources, observers explore the power set of the system (the set of all subsets of the system) and search for some entropy-reducing subset from which an agent and a goal can be defined. They then discard the superfluous degrees of freedom, thereby compressing information. Yet, if the subsystem does indeed reduce entropy, they are still able to make predictions. With successful predictions the world begins to makes sense, so ascribing agency is in a way equivalent to constructing knowledge. In fact, the construction of knowledge can be argued to be the essence of a mind. In the last paragraphs, we have been observing observers. In doing so, we have placed our- selves one step above the hierarchical ladder of observation. We have concluded that purposeful agents do not exist per se, they are a mental construct of observers. This view may be easily accepted when regarding the agents (let us call them "zombies") around us, but is more prob- lematic when it applies to ourselves. In the end, we experience our own purposes in a most irrefutable manner. In the present context, a self with purposes is no more than a compressed representation of an observer. Within a physicalist's point of view, the observer of the self can- not be situated in any other place than in the same brain where the self emerges. Whether the observer coincides with the self, whether it only partially overlaps with it, whether it contains it, or is contained by it, I do not dare to assert. I conjecture, however, that brains create selves following the same principles with which they ascribe agency to external factors, the difference being that the creation of the self involves a vastly larger number of degrees of freedom. Those 9 degrees of freedom, moreover, are typically only accessible to the local subsystem. They in- clude the mental processes of which we have conscious access, encompassing external sensory input, and the detailed state of our body. The latter has been proposed as the base for emotion, and the higher-level neural patterns triggered by such a state, the base of feeling [8]. Within the self, the subject and the object of observation seem to coincide, forming a strange loop. Douglas Hofstadter [9] has suggested that the circular nature of the mind observing itself is essential to the self. I am not sure, however, whether this recursive hypothesis constitutes an actual explanation of the self, or simply a way to bind the two loose ends together and worry no more. It could also be the case that what we perceive as a unitary self is in fact a whole collection of disperse mental processes, inside which multiple observers coexist, although sep- arately unaccessible. In the end, consciousness has been equated with complex and indivisible information processing [10], so accessing subprocesses may not be possible. Dennet however very strongly argues that if such mental subprocesses can be considered multiple observers [11], there is no such thing as a hierarchy, and even less, an ultimate observer. I am afraid I am unable to provide a finished picture of agency when going all the way up to the self. I hope, however, to have built a sensible image of other less intimate agents. The main conclusion of this essay is that the interesting part of agency is the observer. Physics does not make sense, observers make sense of it. Life does not have a meaning, we give it a meaning. Life may not even be fundamentally different from non-life, it may just be a collection of subsystems that appear to have goals. Goal-directed behavior does not exist if we do not define our variables in such a way as to bring goals into existence. Bringing goals into existence is a task that brains perform naturally, because they have evolved to model and predict the future. One fundamental agent that has emerged inside each one of us is the self. The mechanisms behind this process remain unclear, but its evolutionary utility is undisputed. Were we not able to produce meaning, we would not manage to distinguish ourselves as a special part of the cosmos. We would not have a sense of identity, a sense of self-preservation, nor the ability to think. The self is required to enunciate even the most basic statements, all the way up from cogito ergo sum. 10 References [1] Maxwell J C (1908). Theory of heat. London, New York: Longmans, Green. [2] Bennett C H (1982). The thermodynamics of computation -- a review. Int.J.Theoret.Phys. 21 (12): 905940. [3] Landauer RW (1961). Irreversibility and Heat Generation in the Computing Process IBM J.Res. Dev. 5(3): 183 - 191. [4] Fredkin E, Toffoli T (1982). Conservative Logic. Int.J. Theoret.Phys. 21 (3-4): 219253. [5] Dennet DC (1995). Darwin's dangerous idea: Evolution and the meaning of life. New York: Simon & Schuster. [6] Norretranders T (1999). The User Illusion: Cutting Consciousness Down to Size. New York: Viking. [7] MacKay D (2003). Information theory, inference and learning algorithms. New York: Cambridge University Press. [8] Damasio AR (1999). The Feeling of What Happens: Body and Emotion in the Making of Consciousness. San Diego: Harcourt Brace & Co. [9] Hofstadter D (2007). I am a strange loop. New York: Basic Books. [10] Tononi G (2008). Consciousness as Integrated Information: a Provisional Manifesto. Biol. Bull. 215: 216 242. [11] Dennet DC (1991) Consciousness explained. Boston: Little, Brown and Co. 11
1907.08801
2
1907
2019-12-19T05:40:07
Learning spatiotemporal signals using a recurrent spiking network that discretizes time
[ "q-bio.NC", "cs.NE", "nlin.AO", "physics.bio-ph" ]
Learning to produce spatiotemporal sequences is a common task that the brain has to solve. The same neural substrate may be used by the brain to produce different sequential behaviours. The way the brain learns and encodes such tasks remains unknown as current computational models do not typically use realistic biologically-plausible learning. Here, we propose a model where a spiking recurrent network of excitatory and inhibitory biophysical neurons drives a read-out layer: the dynamics of the driver recurrent network is trained to encode time which is then mapped through the read-out neurons to encode another dimension, such as space or a phase. Different spatiotemporal patterns can be learned and encoded through the synaptic weights to the read-out neurons that follow common Hebbian learning rules. We demonstrate that the model is able to learn spatiotemporal dynamics on time scales that are behaviourally relevant and we show that the learned sequences are robustly replayed during a regime of spontaneous activity.
q-bio.NC
q-bio
Learning spatiotemporal signals using a recurrent spiking network that discretizes time Amadeus Maes1, Mauricio Barahona2, and Claudia Clopath1 1Department of Bioengineering, Imperial College London, United Kingdom 2Department of Mathematics, Imperial College London, United Kingdom 9 1 0 2 c e D 9 1 ] . C N o i b - q [ 2 v 1 0 8 8 0 . 7 0 9 1 : v i X r a Abstract Learning to produce spatiotemporal sequences is a common task that the brain has to solve. The same neural substrate may be used by the brain to produce different sequential behaviours. The way the brain learns and encodes such tasks remains unknown as current computational models do not typically use realistic biologically- plausible learning. Here, we propose a model where a spiking recurrent network of excitatory and inhibitory biophysical neurons drives a read-out layer: the dynamics of the driver recurrent network is trained to encode time which is then mapped through the read-out neurons to encode another dimension, such as space or a phase. Different spatiotemporal patterns can be learned and encoded through the synaptic weights to the read-out neurons that follow common Hebbian learning rules. We demonstrate that the model is able to learn spatiotemporal dynamics on time scales that are behaviourally relevant and we show that the learned sequences are robustly replayed during a regime of spontaneous activity. Author summary The brain has the ability to learn flexible behaviours on a wide range of time scales. Previous studies have successfully built spiking network models that learn a variety of computational tasks, yet often the learning involved is not biologically plausible. Here, we investigate a model that uses biological-plausible neurons and learning rules to learn a specific computational task: the learning of spatiotemporal sequences (i.e., the temporal evolution of an observable such as space, frequency or channel index). The model architecture facilitates the learning by separating the temporal information from the other dimension. The time component is encoded into a recurrent network that exhibits sequential dynamics on a behavioural time scale, and this network is then used as an engine to drive the read-out neurons that encode the spatial information (i.e., the second dimension). We demonstrate that the model can learn complex spatiotemporal spiking dynamics, such as the song of a bird, and replay the song robustly spontaneously. Introduction Neuronal networks perform flexible computations on a wide range of time scales. While individual neurons operate on the millisecond time scale, behaviour time scales typically span from a few milliseconds to hun- dreds of milliseconds and longer. Building functional models that bridge this time gap is of increasing in- terest [Abbott et al., 2016], especially now that the activity of many neurons can be recorded simultaneously [Jun et al., 2017, Maass, 2016]. Many tasks and behaviours in neuroscience consist of learning and producing flexible spatiotemporal sequences, e.g. a 2-dimensional pattern with time on the x-axis and any other observable on the y-axis which we denote here in general terms as the "spatial information". For example, songbirds pro- duce their songs through a specialized circuit: neurons in the HVC nucleus burst sparsely at very precise times to drive the robust nucleus of the arcopallium which in its turn drives motor neurons [Hahnloser et al., 2002, Leonardo and Fee, 2005]. For different motor tasks, sequential neuronal activity is recorded in various brain regions [Pastalkova et al., 2008, Itskov et al., 2011, Harvey et al., 2012, Peters et al., 2014, Adler et al., 2019], and while the different tasks involve different sets of muscles, the underlying computation on a more fundamental level might be similar [Rhodes et al., 2004]. Theoretical and computational studies have shown that synaptic weights of recurrent networks can be set appropriately so that dynamics on a wide range of time scales is produced [Litwin-Kumar and Doiron, 2014, Zenke et al., 2015, Tully et al., 2016]. In general, these synaptic weights are engineered to generate a range of interesting dynamics. In slow-switching dynamics, for instance, the wide range of time scales is produced by 1 having stochastic transitions between clusters of neurons [Schaub et al., 2015]. Another example is sequential dynamics, where longer time scales are obtained by clusters of neurons that activate each other in a sequence. This sequential dynamics can emerge by a specific connectivity in the excitatory neurons [Chenkov et al., 2017, Setareh et al., 2018] or in the inhibitory neurons [Billeh and Schaub, 2018]. However, it is unclear how the brain learns these dynamics, as most of the current approaches use non biologically plausible ways to set or "train" the synaptic weights. For example, FORCE training [Sussillo and Abbott, 2009, Laje and Buonomano, 2013, Nicola and Clopath, 2017] or backpropagation through time [Werbos, 1990] use non-local information either in space or in time to update weights. Such information is not available to the synaptic connection, which only has access to the presynaptic and postsynaptic variables at the current time. Here, we propose to learn a spatiotemporal task over biologically relevant time scales using a spiking recurrent network driving a read-out layer where the neurons and synaptic plasticity rules are biologically plausible. Specifically, all synapses are plastic under typical spike-timing dependent Hebbian learning rules [Clopath et al., 2010, Litwin-Kumar and Doiron, 2014]. Our model architecture decomposes the problem into two parts. First, we train a recurrent network to generate a sequential activity which serves as a temporal backbone so that it operates as a 'neuronal clock' driving the downstream learning. The sequential activity is generated by clusters of neurons activated one after the other: as clusters are highly recurrently connected, each cluster undergoes reverberating activity that lasts longer than neural time scale so that the sequential cluster activation is long enough to be behaviourally relevant. This construction allows us to bridge the neural and the behavioural time scales. Second, we use Hebbian learning to encode the target spatiotemporal dynamics in the read-out neurons. In this way, the recurrent network encodes time and the read-out neurons encode 'space'. As discussed above, we use the term 'space' to denote a temporally-dependent observable, be it spatial position, or phase, or a time-dependent frequency, or a more abstract state-space. Similar to the liquid state-machine, where the activity in a recurrent network is linearly read-out by a set of neurons, we can learn different dynamics in parallel in different read-out populations [Jaeger et al., 2007]. We also show that learning in the recurrent network is stable during spontaneous activity and that the model is robust to synaptic failure. Results Model architecture The model consists of two separate modules: a recurrent network and a read-out layer (Fig 1A). Learning happens in two stages. In the first stage, we learn the weights of the recurrent network so that the network exhibits a sequential dynamics. The ensuing recurrent neuronal network (RNN) effectively serves as a temporal backbone driving the learning of the downstream read-out layer. In the second stage, a target sequence is learned in the read-out layer. Architecture: The recurrent network is organized in C clusters of excitatory neurons and a central clus- ter of inhibitory neurons. All excitatory neurons follow adaptive exponential integrate-and-fire dynamics [Brette and Gerstner, 2005] while all inhibitory neurons follow a leaky integrate-and-fire dynamics. The in- hibitory neurons in the RNN prevent pathological dynamics. The aim of this module is to discretize time into C sequential intervals, associated with each of the C clusters. This is achieved by learning the weights of the recurrent network. The neurons in the excitatory clusters then drive read-out neurons through all-to-all feedforward connections. The read-out neurons are not interconnected. The target sequence is learned via the weights between the driver RNN and the read-out neurons. In previous models, the learning schemes are typically not biologically plausible because the Plasticity: plasticity depends on non-local information. Here, however, we use the voltage-based STDP plasticity rule in all the connections between excitatory neurons (Fig 1B). This is paired with weight normalization in the recurrent network (Fig 1C) and weight dependent potentiation in the read-out synapses (Fig 1D). Inhibitory plasticity [Vogels et al., 2011] finds good parameters aiding the sequential dynamics (Fig S5). Learning scheme: During the first stage of learning, all neurons in each cluster receive the same input in a sequential manner. As a result of this learning stage, the recurrent spiking network displays a sequential dynamics of the C clusters of excitatory neurons. Neurons within each cluster spike over a time interval (while all neurons from other clusters are silent), with the activity switching clusters at points t = [t0, t1, ..., tC] so that cluster i is active during time interval [ti−1, ti]. Thus, time is effectively discretized in the RNN. During the second stage of learning, the read-out neurons receive input from a set of excitatory supervisor neurons. The discretization of time enables Hebbian plasticity to form strong connections from the neurons in the relevant time bin to the read-out neurons. For instance, if we want to learn a signal which is 'on' during [ti−1, ti] and 'off' otherwise, a supervisor neuron can activate the read-out neuron during that time interval so 2 Fig 1. Model architecture. (A) The recurrent network consists of both inhibitory (in blue) and excitatory (in red) neurons. The connectivity is sparse in the recurrent network. The temporal backbone is established in the recurrent network after a learning phase. Inset: zoom of recurrent network showing the macroscopic recurrent structure after learning, here for 7 clusters. The excitatory neurons in the recurrent network project all-to-all to the read-out neurons. The read-out neurons are not interconnected. (B) All excitatory to excitatory connections are plastic under the voltage-based STDP rule (see Methods for details). The red lines are spikes of neuron j (top) and neuron i (bottom). When neurons j and i are very active together, they form bidirectional connections strengthening both Wij and Wji. Connections Wij are unidirectionally strengthened when neuron j fires before neuron i. (C) The incoming excitatory weights are L1 normalized in the recurrent network, i.e. the sum of all incoming excitatory weights is kept constant. (D) Potentiation of the plastic read-out synapses is linearly dependent on the weight. This gives weights a soft upper bound. 3 Model architectureRecurrent networkRead-out neurons...ZoomAVoltage-based STDPWWminWmaxALTP0AALTP = A Wmax- WWmax- WminjiWijWjiBidirectionalUnidirectionalWijWjiWijWeight dependent potentiation in read-out synapsesW1W2W1+ W2 = CNormalization in recurrent networkW1W2WBDC that connections from cluster i to the read-out neuron are potentiated through activity (who fires together, wires together). This means that, after learning, the read-out neuron will be activated when cluster i is activated. In general, the read-out layer learns a multivariate signal of time, i.e., the neurons in the read-out layer encode the D different dimensions of the target signal: t → φ(t) = [φ1(t), φ2(t), ..., φD(t)]. A recurrent network that encodes discrete time We give here further details of the first learning stage, where a recurrent network is trained to produce a sequential dynamics. To this end, we initialize the weight matrix so that each synaptic weight between two neurons is non-zero with probability p. The weights that are zero remain zero at all times, i.e. the topology is fixed. We set the initial values of the non-zero weights in the recurrent network such that the dynamics is irregular and asynchronous (i.e., a balanced network, see Methods for details). We stimulate the C clusters with an external input in a sequential manner (Fig 2A): neurons in cluster i each receive external Poisson spike trains (rate of 18 kHz for 10 ms, assuming a large input population). After this, there is a time gap where no clusters receive input (5 ms). This is followed by a stimulation of cluster i + 1. This continues until the last cluster is reached and then it links back to the first cluster (i.e. a circular boundary condition). During the stimulation, neurons in the same cluster fire spikes together strengthening the intra- cluster connections bidirectionally through the voltage-based STDP rule [Clopath et al., 2010, Ko et al., 2013]. Additionally, there is a pre/post pairing between adjacent clusters. Neurons in cluster i + 1 fire after neurons in cluster i. The weights from cluster i to cluster i+1 strengthen unidirectionally (Fig 2B). If the time gap between sequential stimulations is increased during the training phase, so that the gap becomes too long with respect to the STDP time window, then there is no pre/post pairing between clusters and the ensuing dynamics loses its sequential nature and becomes a slow-switching dynamics [Litwin-Kumar and Doiron, 2014, Schaub et al., 2015] (Fig S1). In slow-switching dynamics, clusters of neurons are active over long time scales, but both the length of activation and the switching between clusters is random. This is because the outgoing connections from cluster i to all other clusters are the same in a slow-switching network. To summarize, a connectivity structure emerges through biophysically plausible potentiation by sequentially stimulating the clusters of excitatory neurons in the recurrent network. When the gap between activation intervals is sufficiently small compared to the STDP window, the connectivity structure is such that intra-cluster weights and the weights from successive clusters i → i + 1 are strong. After the synaptic weights have converged, the external sequential input is shut-down and spontaneous dynamics is simulated so that external excitatory Poisson spike trains without spatial or temporal structure drive the RNN. Under such random drive, the sequence of clusters reactivates spontaneously and ensures that both the intra-cluster and the connections from cluster i to cluster i+1 remain strong. In general, the interaction between plasticity, connectivity and spontaneous dynamics can degrade the learned connectivity and lead to unstable dynamics [Morrison et al., 2007]. To test the stability of the learned connectivity, we track the changes in the off-diagonal weights (i.e. the connections from cluster i to cluster i + 1). After the external sequential input is shut-down, we copy the weight matrix and freeze the weights of this copy. We run the dynamics of the recurrent network using the copied frozen weights and apply plastic changes to the original weight matrix. This means that we effectively decouple the plasticity from the dynamics. Indeed, when the dynamics is sequential, the off-diagonal structure is reinforced. When the off-diagonal structure is removed from the frozen copied weight matrix, the dynamics is not sequential anymore. In this case, the off-diagonal structure degrades. We conclude that the connectivity pattern is therefore stable under spontaneous dynamics (Fig S2). We next studied how the spectrum of the recurrent weight matrix is linked to the sequential dynamics. In linear systems, the eigenvalues of the connectivity matrix determine the dynamics of the system. In a nonlinear spiking model, the relationship between connectivity and dynamics is less clear. The connectivity after learning can be seen as a low-dimensional perturbation of a random matrix. Such low-dimensional perturbations create outliers in the spectrum [Tao, 2013] and change the dynamics [Mastrogiuseppe and Ostojic, 2018]. Here, we have carried out a similar spectral analysis to that presented in [Schaub et al., 2015] (see Figs 7 in the Methods and S2 in the Supplementary material). The weight matrix has most of its eigenvalues in a circle in the complex plane (Fig 2B) with eigenvalues associated both with the balanced nature of the network, but, importantly, also with the sequential structure (Fig 2B). As the temporal backbone develops through learning (as seen in Fig S2), it establishes a spectral structure in which the pairs of leading eigenvalues with large real parts have almost constant imaginary parts. A simplified analysis of a reduced weight matrix (where nodes are associated with groups of neurons) shows that the imaginary parts of the dominant eigenvalues depend linearly on the strength of the weights from cluster i to cluster i+1 (see Methods, Fig 7). Hence for this simplified linearised rate model, this results in an oscillatory dynamics where the imaginary part determines the frequency by which the pattern of activation returns due to the periodic excitation pattern. As shown in [Schaub et al., 2015], these properties of the linear system carry over to the nonlinear spiking model, i.e., the imaginary parts of the eigenvalues with large real parts determine the time scales of the sequential activity (Fig S2). 4 Fig 2. Learning a sequential dynamics stably under plasticity. (A) The excitatory neurons receive sequential clustered inputs. Excitatory neurons are grouped in 30 disjoint clusters of 80 neurons each. (7 clusters shown in the cartoon for simplicity) (B) The weight matrix after training (only the first five clusters shown) exhibits the learned connectivity structure, e.g., neurons within cluster 1 are highly interconnected and also project to neurons in cluster 2, same for cluster 2 to cluster 3, etc. The spectrum of the full weight matrix after training shows most eigenvalues in a circle in the complex plane (as in a random graph) with two other eigenvalues signifying the balancing of the network, and a series of dominant eigenvalues in pairs that encode the feedforward embedding. (C) Raster plot of the total network consisting of 2400 excitatory (in red) and 600 inhibitory (in blue) neurons. After learning, the spontaneous dynamics exhibits a stable periodic trajectory 'going around the clock'. The excitatory clusters discretize time (see zoom) and the network has an overall period of about 450 ms. 5 ASequential external input during learningLearning a feedforward structure through the recurrent networkTime123456710 ms5 msTime123456710 ms5 msAfter learning RNN discretizes time under spontaneous random activityCBWeight matrix of RNN after learningDetail of weight matrix(5 first clusters)FeBcnSpectrum(full matrix)Rdc Under spontaneous activity, each cluster is active for about 15 ms, due to the recurrent connectivity within the cluster. A large adaptation current counteracts the recurrent reverberating activity to turn off the activity reliably. Therefore, as each cluster is spontaneously active in a sequence, the sequence length reaches behavioural time scales (Fig 2C). In summary, the network exhibits sequential dynamics, serving as a temporal backbone where time is discretized over behavioural time scales. Learning a non-Markovian sequence After the sequential temporal backbone is learnt via the RNN, we can then learn a spatiotemporal sequence via the read-out neurons. To achieve this, during the second stage of training, the read-out neurons receive additional input from supervisor neurons and from interneurons (Fig 3A). The supervisor neurons receive an external Poisson input with rate modulated by the target sequence to be learned (Fig 3B). As a first example, consider a target sequence composed of states A, B, C activated in the following deterministic order: ABCBA. This is a non-Markovian state sequence because the transition from state B to the next state (A or C) requires knowledge about the previous state [Brea et al., 2013], a non trivial task that requires information to be stored about previous network states, potentially over long time periods. Previous studies have proposed various solutions for this task [Maass and Markram, 2002, Brea et al., 2013]. However, separating the problem of sequence learning in two stages solves this in a natural way. The recurrent network trained in the first stage (Fig 2) is used to encode time. The underlying assumption is that a starting signal activates both the first cluster of the recurrent network and the external input to the supervisor neurons, which activate the read-out neurons. After the training period, the interneurons and supervisor neurons stop firing (Fig 3C) and the target sequence is stored in the read-out weight matrix (Fig 3D). During spontaneous activity, clusters in the RNN reactivate in a sequential manner driving the learned sequence in the read-out neurons. Hence the spike sequence of the read-out neurons is a noisy version of the target signal (Fig 3E). Learning the same target signal several times results in slightly different read-out spike sequences each time (Fig S3). The firing rates of neurons in the read-out corresponds to the target sequence (Fig 3F). In summary, our results show that the model is able to learn simple but non-trivial spatiotemporal signals that are non-Markovian. Learning sequences in parallel We next wondered how multiple spatiotemporal signals can be learned. We hypothesized that, once the temporal backbone is established, multiple spatiotemporal sequences can easily be learned in parallel. As an example, we learn two sequences: ABCBA and DEDED. Here, D and E denote two additional read-out neurons (Fig 4A). We assume that the model observes each sequence alternately for 2 seconds at a time (Fig 4B), although in principle it could also been shown simultaneously. After learning, the target sequences are encoded in the read-out weight matrix (Fig 4C). In a regime of spontaneous dynamics the learned sequences can be replayed (Fig 4D). We conclude that multiple sequences can be learned in parallel. Each separate sequence requires a separate set of read-out neurons. As such, the number of read-out neurons required increases linearly with the number of target sequences. Properties of the model: scaling, robustness and temporal variability We investigate several scaling properties of the network. We first assess how the sequential dynamics in the RNN depends on the cluster size by increasing the number of excitatory neurons in each cluster (NC), preserving the ratio of excitatory to inhibitory neurons (NE/NI). To preserve the magnitude of the currents in the network, the sparseness of the connectivity (p) also varies with NC such that pNC is constant. The same training protocols are used for each network configuration as described in the Methods. In Fig 5A, we show that for a fixed number of clusters C, the mean period of the sequential dynamics exhibited by the RNN is largely independent of cluster size NC. If we fix the number of neurons in the RNN, and in this way change the number of clusters C, the mean period of the sequential dynamics decreases with increasing cluster size NC. We conclude that the sequential dynamics is preserved over a wide range of network configurations. The time scales in the dynamics depend on the number of clusters and network size. Another way to modulate the period of the sequential dynamics is to change the unstructured Poisson input to the RNN during spontaneous dynamics (i.e., after the first stage of learning). When the rate of the external excitatory input is increased/decreased, the mean period of the sequential dynamics in the RNN decreases/increases (Fig 5B). These results suggest that the network could learn even if the supervisor signal changes in length at each presentation, assuming that both the supervisor and external Poisson input are modulated by the same mechanism. We next looked at the robustness of the learning of our model under random perturbations and network size. In this context, we consider the effect of cluster size and the deletion of synapses in the read-out layer 6 Fig 3. Learning a non-Markovian sequence via the read-out neurons. (A) Excitatory neurons in the recurrent network are all-to-all connected to the read-out neurons. The read-out neurons receive additional excitatory input from the supervisor neurons and inhibitory input from interneurons. The supervisor neurons receive spike trains that are drawn from a Poisson process with a rate determined by the target sequence. The read-out synapses are plastic under the voltage-based STDP rule. (B) The rate of the input signal to the supervisor neurons A, B and C. The supervisor sequence is ABCBA where each letter represents a 75 ms external stimulation of 10 kHz of the respective supervisor neuron. (C) After learning, the supervisor input and plasticity are turned off. The read-out neurons are now solely driven by the recurrent network. (D) The read-out weight matrix W RE after 12 seconds of learning. (E) Under spontaneous activity, the spikes of recurrent network (top) and read-out (bottom) neurons. Excitatory neurons in the recurrent network reliably drive sequence replays. (F) The target rate (above) and the rate of the read-out neurons (below) computed using a one sequence replay and normalized to [0, 1]. The spikes of the read-out neurons are convolved with a Gaussian kernel with a width of ∼ 12 ms. 7 Learning a sequence through the read-out neuronsRead-out neuron50150250350CBATarget sequence50150250350Time [ms]CBARead-out neuron rate InterneuronSupervisorneuronTime-discretizing sequential RNNSpontaneous random activity (Poisson)Supervisor input(target)CBATime-discretizing sequential RNNSpontaneous random activity (Poisson)Read-out neuronsReplaying a learned sequenceSupervisor signalRead-out weight matrix after learningRate of read-out neuronsSpontaneous sequence reactivationsA Fig 4. Learning sequences in parallel. (A) The recurrent network projects to two sets of neurons. (B) Two different sequences, ABCBA and DEDED, are learned by alternating between them and presenting each for 2 seconds at a time. (C) The read-out weight matrix after 24 seconds of learning. (D) Raster plot of spontaneous sequence reactivations, where an external inhibitory current is assumed to control which sequence is replayed. 8 0100200300400Time [ms]ABCDESupervisor input rate180016002400Recurrent network neuron indexEDCBARead-out neuron02400.511.52800160024003000Recurrent networkneuron index00.511.52Time [s]ABCDERead-out neuronLearning twosequences in parallel CBATime-discretizing sequential RNNSpontaneous random activity (Poisson)Read-out neuronsAED Fig 5. Scaling, robustness and time variability of the model. (A) Change of the mean period of the sequential dynamics as the number of clusters grows with: (i) the total number of excitatory neurons kept constant (red line); (ii) the total number of neurons increasing with cluster size (blue line). Error bar shows a standard deviation. (B) Dynamics with varying level of external excitatory input for four different cluster sizes and NE = 2400. The external input can modulate the period of the sequential dynamics by ∼ 10%. (C) Recall performance of the learned sequence ABCBA for varying cluster sizes and NE = 30NC under synapse deletion (computed over 20 repeats). The learning time depends on the cluster size: ∆t = 960s/NC. (D) The ABCBA sequence is learned with a network of 120 excitatory neurons connected in one large chain and readout neurons with maximum synaptic read-out strength increased to W AE max = 75 pF . The network is driven by a low external ext = 2.75 kHz). When, at t = 500 ms a single synapse is deleted, the dynamics breaks down and parts input (rEE of the sequence are randomly activated by the external input. Top: spike raster of the excitatory neurons of the RNN. Bottom: spike raster of the read-out neurons. (E) (Left) Histogram of the variability of the period of the sequential activity of the RNN over 79 trials (Right) The standard deviation of the cluster activation time, µt (root mean squared σt, increases as the square root of µt, the mean time of cluster activation: σt = 0.213 error = 0.223 ms). √ 9 Time [s]Deleted read-out synapses per clusterPerformance02040608010012014000.20.40.60.81NC=80NC=120NC=160NC=200Scaling the networkPerformance under synapse deletionSequential dynamics breaks down in synfire chainTemporal variability in sequential activity0200400600Mean time [ms]0246Standard deviation [ms]simulationfit44.254.54.755External input [kHz]250300350400450500550Period [ms]NC= 80NC= 120NC= 160NC= 200External input can modulate the period ABCDEBroken chain after learning. We learn the simple ABCBA sequence (Fig 3) in the read-out neurons using a RNN with a fixed number of clusters C but varying the cluster size NC. The total learning time (∆t) is varied with the cluster size, NC∆t, because smaller clusters learn slower, since smaller clusters need larger read-out synaptic strengths to drive the same read-out neuron. We also eliminate an increasing number of synapses in the read-out layer. Performance is quantified as the number of spikes elicited by the read-out neurons after deletion of read-out synapses normalized by the number of spikes elicited before deletion. Networks with larger clusters show a more robust performance under noise (Fig 5C and Fig S4). These results show that, not surprisingly, larger clusters drive read-out neurons more robustly and learn faster. We then tested the limits of time discretization in our model. To that end, we hardcoded a recurrent network with clusters as small as one neuron. In that extreme case, our network becomes a synfire chain with a single neuron in every layer [Abeles, 1991]. In this case, randomly removing a synapse in the network will break the sequential dynamics (Fig 5D). Hence, although a spatiotemporal signal can be learned in the read-out neurons, the signal is not stable under a perturbation of the synfire chain. In summary, the choice of cluster size is a trade-off between network size on the one hand and robustness on the other hand. Large clusters: (i) require a large network to produce sequential dynamics with the same period; (ii) are less prone to a failure of the sequential dynamics; (iii) can learn a spatiotemporal signal faster; and (iv) drive the read-out neurons more robustly. We have also characterized the variability in the duration of the sequential activity, i.e., the period of the RNN. Since the neural activity does not move through the successive clusters with the same speed in each reactivation, we wondered how the variance in the period of our RNN network compared to Weber's law. Weber's law predicts that the standard deviation of reactions in a timing task grows linearly with time [Gibbon, 1977, Hardy and Buonomano, 2018]. Since our model operates on a time scale that is behaviourally relevant, it is interesting to look at how the variability increases with increasing time. Because time in our RNN is discretized by clusters of neurons that activate each other sequentially, the variability increases over time as in a standard T rather than Markov chain diffusive process. Hence the variability of the duration T is expected to grow as linearly. This is indeed what our network displays (Fig 5E). Here, we scaled the network up and increased the period of the recurrent network by increasing the network size (80 excitatory clusters of 80 neurons each, see Methods for details). By doing so, we can look at how the standard deviation of the duration of the RNN activity grows with time. √ Learning a complex sequence In the non-Markovian target sequence ABCBA, the states have the same duration and the same amplitude (Fig 3B). To test whether we could learn more complex sequences, the model was trained using a spatiotemporal signal with components of varying durations and amplitudes. As an example, we use a 'spatio'-temporal signal consisting of a 600 ms meadowlark song (Fig 6A). The spectrogram of the sound is normalized and used as the time-varying and amplitude-varying rate of the external Poisson input to the supervisor neurons. Each read- out and supervisor neuron encodes a different frequency range, hence in this example our 'space' dimension is frequency. We first trained a RNN of 6400 excitatory neurons (Fig 5E, see Methods) in order to discretize the time interval spanning the full duration of the song. We then trained the read-out layer. The learned read-out weight matrix reflects the structure of the target sequence (Fig 6B). Under spontaneous activity, the supervisor neurons and interneurons stop firing and the recurrent network drives song replays (Fig 6C), and the learned spatiotemporal signal broadly follows the target sequence (Fig 6A). The model performs worse when the target dynamics has time-variations that are faster than or of the same order as the time discretization in the RNN. Thus, we conclude that the model can learn interesting spiking dynamics up to a resolution of time features limited by the time discretization in the recurrent network. Discussion We have proposed here a neuronal network architecture based on biophysically plausible neurons and plasticity rules in order to learn spatiotemporal signals. The architecture is formed by two modules. The first module pro- vides a temporal backbone that discretizes time, implemented by a recurrent network where excitatory neurons are trained into clusters that become sequentially active due to strong inter-cluster and cluster i to cluster i + 1 weights. All of the excitatory clusters are connected to a central cluster of inhibitory neurons. As previously shown for randomly switching clustered dynamics [Litwin-Kumar and Doiron, 2014], the ongoing spontaneous activity does not degrade the connectivity patterns: the set of plasticity rules and sequential dynamics reinforce each other. This stable sequential dynamics provides a downstream linear decoder with the possibility to read out time at behavioural time scales. The second module is a set of read-out neurons that encode another dimen- sion of a signal, which we generically denote as 'space' but can correspond to any time-varying observable, e.g., 10 Fig 6. Learning a complex sequence. (A) Target sequence (top). The amplitude shows the rate of the Poisson input to the supervisor neurons and is normalized between 0 and 10 kHz. Rate of read-out neurons for one sample reactivation after learning 6 seconds (bottom). 45 read-out neurons encode the different frequencies in the song. Neuron i encodes a frequency interval of [684 + 171i, 855 + 171i]Hz. (B) The read-out weight matrix after learning 6 seconds. (C) Sequence replays showing the spike trains of both the recurrent network neurons (top, excitatory neurons in red and inhibitory neurons in blue), and the read-out neurons (bottom). spatial coordinates, frequency, discrete states, etc. The read-out neurons learn spike sequences in a supervised manner, and the supervisor sequence is encoded into the read-out weight matrix. Bringing together elements from different studies [Brea et al., 2013, Nicola and Clopath, 2017, Gilra and Gerstner, 2017], our model ex- ploits a clock-like dynamics encoded in the RNN to learn a mapping to read-out neurons so as to perform the computational task of learning and replaying spatiotemporal sequences. We illustrated the application of our scheme on a simple non-Markovian state transition sequence, a combination of two such simple sequences, and a time series from bird singing with more complex dynamics. Other studies have focused on the classification of spatiotemporal signals. The tempotron classifies a spa- tiotemporal pattern by either producing a spike or not [Gütig and Sompolinsky, 2006]. More recent studies on sequential working memory propose similar model architectures that enable the use of Hebbian plasticity [Manohar et al., 2019]. For example, spatiotemporal input patterns can be encoded in a set of feedforward synapses using STDP-type rules [Park et al., 2017, Lee et al., 2019]. Combining these approaches with our model might be an interesting line of future research. The dynamics of the recurrent network spans three time scales: (i) individual neurons fire at the millisecond time scale; (ii) clusters of neurons fire at the "tick of the clock", τc, i.e., the time scale that determines the time discretization of our temporal backbone; and (iii) the slowest time scale is at the level of the entire network, i.e. the period of the sequential activity, τp, achieved over the cascade of sequential activations of the clusters (see Fig 5A). The time scales τc and τp are dependent on several model parameters: the cluster and network size, the average connection strengths within the clusters, and adaptation. Smaller cluster sizes lead to a smaller τc when the network size is fixed and conversely τp increases with network size when the cluster size is fixed. The recurrent network is the "engine" that, once established, drives read-out dynamics. Our model can learn different read-out synapses in parallel (Fig 4) and is robust to synapse failure (Fig 5C). This robustness is a con- sequence of the clustered organization of the recurrent network. Previously proposed models are also robust to similar levels of noise [Nicola and Clopath, 2017, Nicola and Clopath, 2019]. While an exact comparison is hard to draw, we have shown that it is possible to retain robustness while moving towards a more biological learning rule. The development of a clustered organization in the RNN allows a large drive for the read-out neurons while keeping the individual synaptic strengths reasonably small. If the clusters become small, larger read-out synap- tic strengths are required, and the dynamics become less robust. Indeed, the sequential dynamics is especially fragile in the limit where every cluster has exactly one neuron. Furthermore, we show that learning is faster with more neurons per cluster since relatively small changes in the synapses are sufficient to learn the target. This is consistent with the intuitive idea that some redundancy in the network can lead to an increased learning speed [Raman et al., 2019]. In its current form, the target pattern needs to be presented repeatedly to the net- 11 B Read-out weight matrixC Spontaneous sequence reactivations12004006004525Targetsequence545255452554525451200400600Time [ms]Read-outneuron rate6400RNN neuron indexRead-outneuron index 320018000RNNneuron index00.511.52Read-outneuron indexTime [s]00.511.52A Target and learned sequence4000 work and does not support one-shot learning through a single supervisor presentation. Although increasing the cluster size NC can reduce the number of presentations, the size of the clusters would need to be impractically large for one-shot learning. Alternatively, increasing the learning rate of the plastic read-out weights could be a way to reduce the number of target presentations to just one [Nicola and Clopath, 2019]. However, it is unclear at present whether such high learning rates are supported by experiments [Sjöström et al., 2001]. Taken together, our numerical simulations suggest ways to scale our network. An optimal network con- figuration can be chosen given the temporal length of the target sequence, and requirements on the temporal precision and robustness. For example, we have shown that a network with NE = 6400 and NC = 200 can be chosen for a 400 ms target sequence that can be learned fast with good temporal precision and a robust replay. If the network configuration and size are fixed, this severely constrains the sequences that can be learned and how they are replayed. In this paper, we use local Hebbian learning to produce a sequential dynamics in the recurrent network. This is in contrast with previous studies, where often a recursive least squares method is used to train the weights of the recurrent network [Rajan et al., 2016, Hardy and Buonomano, 2018]. Hardcoding a weight structure into the recurrent network has been shown to result in a similar sequential dynamics [Chenkov et al., 2017, Setareh et al., 2018, Spreizer et al., 2019]. Studies that do incorporate realistic plasticity rules are mostly lim- ited to purely feedforward synfire chains [Fiete et al., 2010, Waddington et al., 2012, Zheng and Triesch, 2014]. Those studies focus on the generation of a dynamics that is sequential. In this regard, the contribution of our work is to use the sequential dynamics as a key element to learning spatiotemporal spiking patterns. The ubiquity of sequential dynamics in various brain regions [Ikegaya et al., 2004, Jin et al., 2009, Mackevicius et al., 2019] and the architecture of the songbird system [Fee and Scharff, 2010] were an inspiration for the proposed sepa- ration of temporal and spatial information in our setup. As we have shown, this separation enables the use of a local Hebbian plasticity rule in the read-out synapses. Our model would therefore predict that perturbing the sequential activity should lead to learning impairments. Further perturbation experiments can test this idea and shed light on the mechanisms of sequential learning [Hemberger et al., 2019]. Previous studies have discussed whether sequences are learned and executed serially or hierarchically [Lashley, 1951]. Our recurrent network has a serial organization. When the sequential activity breaks down halfway, the re- maining clusters are not activated further. A hierarchical structure would avoid such complete breakdowns at the cost of a more complicated hardware to control the system. Sequences that are chunked in sub- sequences can be learned separately and chained together. When there are errors in early sub-sequences this will less likely affect the later sub-sequences. A hierarchical organization might also improve the ca- pacity of the network. In our proposed serial organization, the number of spatiotemporal patterns that can be stored is equal to the number of read-out neurons. A hierarchical system could be one way to extract general patterns and reduce the number of necessary read-out neurons. Evidence for hierarchical structures is found throughout the literature [Sakai et al., 2003, Glaze and Troyer, 2006, Okubo et al., 2015]. The basal ganglia is for example thought to play an important role in shaping and controlling action se- quences [Tanji, 2001, Jin and Costa, 2015, Geddes et al., 2018]. Another reason why a hierarchical organization seems beneficial is inherent to the sequential dynamics. The time-variability of the sequential activity grows by t (see Fig 5E)). While on a time scale of a few hundreds of milliseconds, this does not pose a approximately problem, for longer target sequences this variability would exceed the plasticity time constants. The presented model could thus serve as an elementary building block of a more complex hierarchy. In summary, we have demonstrated that a clustered network organization can be a powerful substrate for learning, moving biological learning systems closer to machine learning performance. Specifically, the model dissociates temporal and spatial information and therefore can make use of Hebbian learning to learn spa- tiotemporal sequences over behavioural time scales. More general, the backbone as a clustered connectivity might encode any variable x and enable downstream read-out neurons to learn and compute any function of this variable, φ(x). √ Methods Neuron and synapse models Excitatory neurons are modelled with the adaptive exponential integrate-and-fire model [Brette and Gerstner, 2005]. A classical integrate-and-fire model is used for the inhibitory neurons. All excitatory to excitatory recurrent synapses are plastic under the voltage-based STDP rule [Clopath et al., 2010]. This enables the creation of neu- ronal clusters and a feedforward structure. Normalization and weight bounds are used to introduce competition and keep the recurrent network stable. Synapses from inhibitory to excitatory neurons in the recurrent network are also plastic under a local plasticity rule [Vogels et al., 2011]. In general, it prevents runaway dynamics and allows for an automatic search of good parameters (Fig S5). The connections from the recurrent network to the read-out neurons are plastic under the same voltage-based STDP rule. However, potentiation of read-out 12 Table 1: Initialization of network 0 Constant Value N E 2400 N I 600 0.2 p 2.83 pF wEE 1.96 pF wIE 62.87 pF wEI 0 20.91 pF wII 0 pF wRE 200 pF wRS 200 pF wRH 200 pF wHR 0 Description Number of recurrent E neurons Number of recurrent I neurons Recurrent network connection probability Initial E to E synaptic strength E to I synaptic strength Initial I to E synaptic strength I to I synaptic strength Initial E to R synaptic strength S to R synaptic strength H to R synaptic strength R to H synaptic strength synapses is linearly dependent on the strength of the synapses. There is no normalization here to allow a contin- uous weight distribution. The dynamics was chosen based on previous models, with parameters for the dynamics and plasticity to a large extent conserved [Litwin-Kumar and Doiron, 2014]. More simple integrate-and-fire dy- namics should lead to the same qualitative results, given that the parameters are appropriately changed (data not shown). Network dynamics Recurrent network A network with N E excitatory (E) and N I inhibitory (I) neurons is homogeneously recurrently connected with connection probability p. Our network is balanced in terms of inhibition and excita- tion, so that it displays irregular and asynchronous spiking. This is signalled by the coefficient of variation (CV) of the inter-spike intervals of the neurons being CV ∼ 1, thus indicating Poisson-like spiking [Brunel, 2000]. In our construction, we initialise the weights of the network near the balanced state by scaling the weights of the balanced RNN in Ref. [Litwin-Kumar and Doiron, 2014] by the square root of the relative network size. We then verify that the scaled parameters indeed lead to irregular dynamics. The spiking dynamics is slightly more regular on average, with a mean CV ∼ 0.8 for excitatory neurons and a mean CV ∼ 0.9 for inhibitory neurons. Read-out neurons The N E excitatory neurons from the recurrent network are all-to-all connected to N R excitatory read-out (R) neurons. This weight matrix is denoted by W RE and it is where the learned sequence is stored. To help learning, there are two additional types of neurons in the read-out network. During learning, the read-out neurons receive supervisory input from N R excitatory supervisor (S) neurons. The connection from supervisor neurons to read-out neurons is one-to-one and fixed, wRS. Also during learning, N R interneurons (H) are one-to-one and bidirectionally connected to the read-out neurons with fixed connection strengths, wRH and wHR (see Table 1 for the recurrent network and read-out parameters). The E to E, I to E and the E to R connections are plastic. Membrane potential dynamics There are two different regimes, one for each part of the model. Excitatory neurons in the recurrent network have a high adaptation current while excitatory neurons in the read-out network have no adaptation. This is to allow for a wide range of firing rates in the read-out network, while spiking is more restricted in the recurrent network. Differences in the refractory period are there for the same reason, but are not crucial. The membrane potential of the excitatory neurons (V E) in the recurrent network has the following dynamics: (cid:18) (cid:18) V E − V E T (cid:19)(cid:19) dV E dt = 1 τ E L − V E + ∆E EE T exp ∆E T + gEE EE − V E C + gEI EI − V E C − aE C (1) where τ E is the membrane time constant, EE T is the slope of the exponential, C is the capacitance, gEE, gEI are synaptic input from excitatory and inhibitory neurons respectively and EE, EI are the excitatory and inhibitory reversal potentials respectively. When the membrane potential diverges and exceeds 20 mV, the neuron fires a spike and the membrane potential is reset to Vr. This reset potential is the same for all neurons in the model. There is an absolute refractory period of τabs. The parameter V E T is adaptive for excitatory neurons and set to VT + AT after a spike, relaxing back to VT with time constant τT : L is the reversal potential, ∆E τT dV E T dt = VT − V E T . 13 (2) The adaptation current aE for recurrent excitatory neurons follows: τa daE dt = −aE. (3) where τa is the time constant for the adaptation current (see also Fig S6). The adaptation current is increased with a constant β when the neuron spikes. The membrane potential of the read-out (V R) neurons has no adaptation current: dV R dt = 1 τ E L − V R + ∆E EE T exp + gRE EE − V R C + gRS EE − V R C + gRH EI − V R C (4) (cid:18) V R − V R T (cid:19)(cid:19) ∆E T (cid:18) L , ∆E where τ E, EE T , EE, EI and C are as defined before. gRE is the excitatory input from the recurrent network. gRS is the excitatory input from the supervisor neuron (supervisor input only non-zero during learning, when the target sequence is repeatedly presented). gRH is the inhibitory input from the interneuron (only non-zero during learning, to have a gradual learning in the read-out synapses). The absolute refractory period is τabsR. The threshold V R T , with the same parameters. The membrane potential of the supervisor neurons (V S) has no inhibitory input and no adaptation current: T follows the same dynamics as V E dV S dt = 1 τ E L − V S + ∆E EE T exp + gSE EE − V S C (5) (cid:18) V S − V S T (cid:19)(cid:19) ∆E T (cid:18) where the constant parameters are defined as before and gSE is the external excitatory input from the target sequence. The absolute refractory period is τabsS. The threshold V S T follows again the same dynamics as V E T , with the same parameters. The membrane potential of the inhibitory neurons (V I) in the recurrent network has the following dynamics: dV I dt = L − V I EI τ I + gIE EE − V I C + gII EI − V I C . (6) L is the inhibitory reversal potential and EE, EI are where τ I is the inhibitory membrane time constant, EI the excitatory and inhibitory resting potentials respectively. gEE and gEI are synaptic input from recurrent excitatory and inhibitory neurons respectively. Inhibitory neurons spike when the membrane potential crosses the threshold VT , which is non-adaptive. After this, there is an absolute refractory period of τabs. There is no adaptation current. The membrane potential of the interneurons (V H) follow the same dynamics and has the same parameters, but there is no inhibitory input: dV H dt = L − V H EI τ I + gHE EE − V H C (7) where the excitatory input gHE comes from both the read-out neuron it is attached to and external input. After the threshold VT is crossed, the interneuron spikes and an absolute refractory period of τabsH follows. The interneurons inhibit the read-out neurons stronger when they receive strong inputs from the read-out neurons. This slows the potentiation of the read-out synapses down and keeps the synapses from potentiating exponentially (see Table 2 for the parameters of the membrane dynamics). Synaptic dynamics The synaptic conductance of a neuron i is time dependent, it is a convolution of a kernel with the total input to the neuron i: (t) = K Y (t) ∗ gXY i ext sX i,ext + W XY ij sY j (t) (8) W X (cid:88) j  . where X and Y denote two different neuron types in the model (E, I, R, S or H). K is the difference of exponentials kernel: K Y (t) = d − e−t/τ Y e−t/τ Y d − τ Y τ Y r r , with a decay time τd and a rise time τr dependent only on whether the neuron is excitatory or inhibitory. There is no external inhibitory input to the supervisor and inter- neurons. During spontaneous activity, there is no external inhibitory input to the recurrent network and a fixed rate. The external input to the interneurons has a fixed rate during learning as well. The external input to the supervisor neurons is dependent on the specific learning task. There is no external input to the read-out neurons. The externally incoming spike trains sX ext are generated from a Poisson process with rates rX ext. The externally generated spike trains enter the network through synapses W X ext (see Table 3 for the parameters of the synaptic dynamics). 14 Table 2: Neuronal membrane dynamics parameters Description E membrane potential time constant I membrane potential time constant Refractory period of E and I neurons R neurons refractory period S neurons refractory period H neurons refractory period excitatory reversal potential inhibitory reversal potential excitatory resting potential inhibitory resting potential Constant Value 20 ms τE 20 ms τI 5 ms τabs 1 ms τabsR 1 ms τabsS 1 ms τabsH 0 mV EE −75 mV EI −70 mV EE −62 mV L EI −60 mV Reset potential (for all neurons the same) L Vr 300 pF C 2 mV ∆E T 30 ms τT −52 mV Membrane potential threshold VT 10 mV AT 100 ms τa 1000 pA Adaptation current increase constant of recurrent network neurons β Capacitance Exponential slope Adaptive threshold time constant Adaptive threshold increase constant Adaptation current time constant Table 3: Synaptic dynamics parameters Description E decay time constant E rise time constant I rise time constant I rise time constant External input synaptic strength to E neurons Rate of external input to E neurons External input synaptic strength to I neurons Constant Value 6 ms τ E d 1 ms τ E r 2 ms τ I d 0.5 ms τ I r 1.6 pF W E ext 4.5 kHz rE ext 1.52 pF W I 2.25 kHz Rate of external input to I neurons rI ext 1.6 pF W S ext 1.6 pF W H ext 1.0 kHz rH ext External input synaptic strength to S neurons External input synaptic strength to H neurons Rate of external input to H neurons ext 15 Table 4: Excitatory plasticity parameters Constant Value ALT D A θLT D θLT P τu 0.0014 pA mV −2 0.0008 pA mV −1 −70 mV −49 mV 10 ms τv τxEE τxRE W EE min W EE max W RE min W RE max 7 ms 3.5 ms 5 ms 1.45 pF 32.68 pF 0 pF 25 pF Description LTD amplitude LTP amplitude (in RNN: ALT P = A) LTD threshold LTP threshold Time constant of low pass filtered postsynaptic membrane potential (LTD) Time constant of low pass filtered postsynaptic membrane potential (LTP) Time constant of low pass filtered presynaptic spike train in recurrent network Time constant of low pass filtered presynaptic spike train for read-out synapses Minimum E to E weight Maximum E to E weight Minimum E to R weight Maximum E to R weight Plasticity Excitatory plasticity The voltage-based STDP rule is used [Clopath et al., 2010]. The synaptic weight from excitatory neuron j to excitatory neuron i is changed according to the following differential equation: = −ALT D sj(t) R(cid:0)ui(t) − θLT D (cid:1) + ALT P xj(t) R(cid:0)Vi(t) − θLT P (cid:1) R(cid:0)vi(t) − θLT D (cid:1). dWij dt (9) Here, ALT D and ALT P are the amplitude of depression and potentiation respectively. θLT D and θLT P are the voltage thresholds to recruit depression and potentiation respectively, as R(.) denotes the linear-rectifying function (R(x) = 0 if x < 0 and else R(x) = x). Vi is the postsynaptic membrane potential, ui and vi are low-pass filtered versions of Vi, with respectively time constants τu and τv (see also Fig S6): τu τv dui dt dvi dt = Vi − ui = Vi − vi where sj is the presynaptic spike train and xj is the low-pass filtered version of sj with time constant τx: τx dxj dt = sj − xj. (10) (11) (12) a neuron constant: (cid:80) j Wij = K. E to E weights have a lower and upper bound [W EE Here the time constant τx is dependent on whether learning happens inside (E to E) or outside (E to R) the recurrent network. sj(t) = 1 if neuron j spikes at time t and zero otherwise. Competition between synapses in the recurrent network is enforced by a hard L1 normalization every 20 ms, keeping the sum of all weights onto max]. The minimum and maximum strengths are important parameters and determine the position of the dominant eigenvalues of W . Potentiation of the read-out synapses is weight dependent. Assuming that stronger synapses are harder to potentiate [Debanne et al., 1999], ALT P reduces linearly with W RE: max − W RE W RE max − W RE W RE min, W EE ALT P = A (13) . min The maximum LTP amplitude A is reached when W RE = W RE plasticity rule). min (see Table 4 for the parameters of the excitatory Inhibitory plasticity Inhibitory plasticity acts as a homeostatic mechanism, previously shown to prevent runaway dynamics [Vogels et al., 2011, Litwin-Kumar and Doiron, 2014, Zenke et al., 2015]. Here, it allows to automatically find good parameters (see also Fig S5). Excitatory neurons that fire with a higher frequency will receive more inhibition. The I to E weights are changed when the presynaptic inhibitory neuron or the postsynaptic excitatory neuron fires [Vogels et al., 2011]: i (t) − 2r0τy j (t) + Ainh yI (cid:1) sI (cid:0)yE j (t) sE = Ainh dWij (14) i (t) dt 16 Table 5: Inhibitory plasticity parameters Description Constant Value Ainh r0 τy W EI min W EI max 10−5 AHz Amplitude of inhibitory plasticity 3 Hz 20 ms 48.7 pF 243 pF Target firing rate Time constant of low pass filtered spike train Minimum I to E weight Maximum I to E weight Table 6: Parameters for the large recurrent network (all the other parameters are the same as the smaller network) 0 Constant Value N E 6400 N I 1600 1.73 pF wEE 1.20 pF wIE 40 pF wEI 0 12.80 pF wII 1.27 pF W EE min 30.5 pF W EE max 40 pF W EI min 200 pF W EI max 15 pF W RE max Description Number of recurrent E neurons Number of recurrent I neurons baseline E to E synaptic strength E to I synaptic strength Initial I to E synaptic strength I to I synaptic strength Minimum E to E weight Maximum E to E weight Minimum I to E weight Maximum I to E weight Maximum E to R weight where r0 is a constant target rate for the postsynaptic excitatory neuron. sE and sI are the spike trains of the postsynaptic E and presynaptic I neuron respectively. The spike trains are low pass filtered with time constant τy to obtain yE and yI (as in equation 12). Table 5 shows parameter values for the inhibitory plasticity rule. The I to E synapses have a lower and upper bound [W IE min, W IE max]. Learning protocol Learning happens in two stages. First a sequential dynamics is learned in the RNN. Once this temporal backbone is established connections to read-out neurons can be learned. Read-out neurons are not interconnected and can learn in parallel. Recurrent network The network is divided in 30 disjoint clusters of 80 neurons. The clusters are sequentially stimulated for a time duration of 60 minutes by a large external current where externally incoming spikes are drawn from a Poisson process with rate 18 kHz. This high input rate does not originate from a single external neuron but rather assumes a large external input population. Each cluster is stimulated for 10 ms and in between cluster stimulations there are 5 ms gaps (see also Fig S6 for different gaps). During excitatory stimulation of a cluster, all other clusters receive an external inhibitory input with rate 4.5 kHz and external input weight ext = 2.4 pF . There is a periodic boundary condition, i.e. after the last cluster is activated, the first cluster W I is activated again. After the sequential stimulation, the network is spontaneously active for 60 minutes. The connectivity stabilizes during the spontaneous dynamics. Learning in scaled versions of this network happens in exactly the same way (Fig 5A). The recurrent weight matrix of the large network (80 clusters of 80 neurons, Figs 5E and 6) is learned using the same protocol. The recurrent weight matrix reaches a stable structure after three hours of sequential stimulation followed by three hours of spontaneous dynamics. Parameters that change for the scaled up version are summarized in Table 6. For randomly switching dynamics, a similar protocol is followed (Fig S1). The weight matrix used to plot the spectrum of the recurrent network in Figs 2 and S2 is: (cid:18) W EE W IE −W EI −W II (cid:19) W = . Read-out network During learning of the read-out synapses, external input drives the supervisor and inter- neurons. The rate of the external Poisson input to the supervisor neurons reflects the sequence that has to be learned. The rate is normalized to be between 0 kHz and 10 kHz. During learning, W RE changes. After learning, the external input to the supervisor and inter- neurons is turned off and both stop firing. The read-out neurons are now solely driven by the recurrent network. Plasticity is frozen in the read-out synapses after learning. With plasticity on during spontaneous dynamics, the read-out synapses would continue to potentiate 17 because of the coactivation of clusters in the recurrent network and read-out neurons. This would lead to read-out synapses that are all saturated at the upper weight bound. Simulations The code used for the training and simulation of the recurrent network is built on top of the code from [Litwin-Kumar and Doiron, 2014] in Julia. The code used for learning spatiotemporal sequences using read-out neurons is written in Matlab. Forward Euler discretization with a time step of 0.1 ms is used. The code is available for the reviewers on ModelDB (http://modeldb.yale.edu/257609) with access code 'SSRSNT'. The code will be made public after publication. Linear rate model: spectral analysis A linear rate model can give insight into the dynamics of a large nonlinear structured spiking network [Schaub et al., 2015]. The dynamics of a simplified rate model with the same feedforward structure as in the RNN is as follows: = −x + Ax + ξ dx dt (15) where x is a multidimensional variable consisting of the rates of all excitatory and inhibitory clusters, A is the weight matrix, and ξ is white noise. The matrix A is a coarse-grained version of the weight matrix of the recurrent network in Fig 2B averaged over each cluster. In order to obtain analytical expressions, we consider a network with 3 excitatory clusters and 1 inhibitory cluster. The connectivity of this model can be parametrized as: where δ > 0, w = δ +  + 1,  > 1 guarantees sequential dynamics, and k > 1 guarantees a balanced network. The Schur decomposition A = U T U T gives eigenvalues and Schur vectors ui: (16) (17) (18)   −kw 1 −kw δ −kw 3 −kw A = √ w  1 w 3 1 δ  w 3  δ (cid:1) 3w(cid:0)k + 1  , u3 =  1 0 0 0 1 3 1−3 0 0 0 0 δ − +1 √ 2 ( − 1) 3 2 3 − √ 2 ( − 1) δ − +1 2 −1  , u4 = 0 1 0 1√ 2 1√ 6  −1  . 2−1 0 −w(k − 1)  , u2 = 0 0 0 2 √ 1 3 T = 1 1 1 1 u1 = 1 2 The first mode u1 decays fast and uniformly over the four neuronal groups. The second mode u2 decays more slowly, and indicates the interplay between excitatory groups (x1, x2, x3) and the inhibitory group x4. The eigenspace associated with the pair {u3, u4} has complex conjugate eigenvalues and is localized on the three excitatory groups. An increase of activity in one excitatory group is coupled with decreased activities in the other groups. If the real part δ− +1 2 < 1 then these modes are linearly stable but if the real part becomes closer to one means a slower decay of this mode. Importantly, the imaginary part of the eigenvalues is ±√ 3( − 1)/2; hence it grows linearly with the strength of the feedforward structure ( − 1) (Fig 7A). This leads to oscillatory behavior, which determines the time scale of the sequential switching. 18 Fig 7. Spectral analysis of reduced linear model. (A) Cartoon of a simplified linearised rate model with three nodes x1, x2, x3 corresponding to three clusters of excitatory neurons with recurrent strength δ connected to a central cluster of inhibitory neurons x4. The cyclic connections are stronger clockwise than anticlockwise since  > 1. (B) The spectrum shows a conjugate complex eigenvalue pair with large real part (2δ −  − 1)/2 and an imaginary part ±√ 3( − 1)/2 which grows linearly with the asymmetry of the clockwise/anticlockwise strength ( − 1). This pair of eigenvalues dominates the dynamics as their real parts are close to 1 and leads to the periodic behaviour corresponding to propagation around the cycle x1 → x2 → x3 → x1 . . .. References [Abbott et al., 2016] Abbott, L. F., DePasquale, B., and Memmesheimer, R. M. (2016). Building functional networks of spiking model neurons. Nature Neuroscience, 19(3):350 -- 355. [Abeles, 1991] Abeles, M. (1991). Corticonics: Neural circuits of the cerebral cortex. Cambridge University Press. [Adler et al., 2019] Adler, A., Zhao, R., Shin, M. E., Yasuda, R., and Gan, W.-B. (2019). Somatostatin- Expressing Interneurons Enable and Maintain Learning-Dependent Sequential Activation of Pyramidal Neu- rons. Neuron, 102(1):202 -- 216.e7. [Billeh and Schaub, 2018] Billeh, Y. N. and Schaub, M. T. (2018). Feedforward architectures driven by in- hibitory interactions. Journal of Computational Neuroscience, 44(1):63 -- 74. [Brea et al., 2013] Brea, J., Senn, W., and Pfister, J.-P. (2013). Matching Recall and Storage in Sequence Learning with Spiking Neural Networks. Journal of Neuroscience, 33(23):9565 -- 9575. [Brette and Gerstner, 2005] Brette, R. and Gerstner, W. (2005). Adaptive exponential integrate-and-fire model as an effective description of neuronal activity. Journal of Neurophysiology, 94(5):3637 -- 3642. [Brunel, 2000] Brunel, N. (2000). Dynamics of Sparsely Connected Networks of Excitatory and Inhibitory Spiking Neurons. Journal of Computational Neuroscience, 8(3):183 -- 208. [Chenkov et al., 2017] Chenkov, N., Sprekeler, H., and Kempter, R. (2017). Memory replay in balanced recur- rent networks. PLoS Computational Biology, 13(1):e1005359. [Clopath et al., 2010] Clopath, C., Büsing, L., Vasilaki, E., and Gerstner, W. (2010). Connectivity reflects coding: A model of voltage-based STDP with homeostasis. Nature Neuroscience, 13(3):344 -- 352. [Debanne et al., 1999] Debanne, D., Gähwiler, B. H., and Thompson, S. M. (1999). Heterogeneity of Synaptic Plasticity at Unitary CA3 -- CA1 and CA3 -- CA3 Connections in Rat Hippocampal Slice Cultures. The Journal of Neuroscience, 19(24):10664 -- 10671. [Fee and Scharff, 2010] Fee, M. S. and Scharff, C. (2010). The Songbird as a Model for the Generation of Complex Behaviors. ILAR journal, 51(4):362 -- 377. [Fiete et al., 2010] Fiete, I. R., Senn, W., Wang, C. Z., and Hahnloser, R. H. (2010). Spike-Time-Dependent Plasticity and Heterosynaptic Competition Organize Networks to Produce Long Scale-Free Sequences of Neural Activity. Neuron, 65(4):563 -- 576. [Geddes et al., 2018] Geddes, C. E., Li, H., and Jin, X. (2018). Optogenetic Editing Reveals the Hierarchical Organization of Learned Action Sequences. Cell, 174(1):32 -- 43.e15. 19 x4A Cartoon of linear rate modelB Eigenvalues of linear rate model-kww/31-101Real axis-0.0500.05Imaginary axisx1x2x33 [Gibbon, 1977] Gibbon, J. (1977). Scalar expectancy theory and Weber's law in animal timing. Psychological Review, 84(3):279 -- 325. [Gilra and Gerstner, 2017] Gilra, A. and Gerstner, W. (2017). Predicting non-linear dynamics by stable local learning in a recurrent spiking neural network. Elife, 6(e28295):1 -- 38. [Glaze and Troyer, 2006] Glaze, C. M. and Troyer, T. W. (2006). Temporal Structure in Zebra Finch Song: Implications for Motor Coding. Journal of Neuroscience, 26(3):991 -- 1005. [Gütig and Sompolinsky, 2006] Gütig, R. and Sompolinsky, H. (2006). The tempotron: A neuron that learns spike timing-based decisions. Nature Neuroscience, 9(3):420 -- 428. [Hahnloser et al., 2002] Hahnloser, R. H., Kozhevnikov, A. A., and Fee, M. S. (2002). An ultra-sparse code underlies the generation of neural sequences in a songbird. Nature, 419(6902):65 -- 70. [Hardy and Buonomano, 2018] Hardy, N. F. and Buonomano, D. (2018). Encoding Time in Feedforward Tra- jectories of a Recurrent Neural Network Model. Neural Computation, 30(2):378 -- 396. [Harvey et al., 2012] Harvey, C. D., Coen, P., and Tank, D. W. (2012). Choice-specific sequences in parietal cortex during a virtual-navigation decision task. Nature, 484(7392):62 -- 68. [Hemberger et al., 2019] Hemberger, M., Shein-Idelson, M., Pammer, L., and Laurent, G. (2019). Reliable Sequential Activation of Neural Assemblies by Single Pyramidal Cells in a Three-Layered Cortex. Neuron, 104(2):353 -- 369.e5. [Ikegaya et al., 2004] Ikegaya, Y., Aaron, G., Cossart, R., Aronov, D., Lampl, I., Ferster, D., and Yuste, R. (2004). Synfire Chains and Cortical Songs: Temporal Modules of Cortical Activity. Science, 304(5670):559 -- 564. [Itskov et al., 2011] Itskov, V., Curto, C., Pastalkova, E., and Buzsaki, G. (2011). Cell Assembly Sequences Arising from Spike Threshold Adaptation Keep Track of Time in the Hippocampus. Journal of Neuroscience, 31(8):2828 -- 2834. [Jaeger et al., 2007] Jaeger, H., Maass, W., and Principe, J. (2007). Special issue on echo state networks and liquid state machines. Neural Networks, 20(3):287 -- 289. [Jin et al., 2009] Jin, D. Z., Fujii, N., and Graybiel, A. M. (2009). Neural representation of time in cortico- basal ganglia circuits. Proceedings of the National Academy of Sciences of the United States of America, 106(45):19156 -- 19161. [Jin and Costa, 2015] Jin, X. and Costa, R. M. (2015). Shaping action sequences in basal ganglia circuits. Current Opinion in Neurobiology, 33:188 -- 196. [Jun et al., 2017] Jun, J. J., Steinmetz, N. A., Siegle, J. H., Denman, D. J., Bauza, M., Barbarits, B., Lee, A. K., Anastassiou, C. A., Andrei, A., Aydin, Ç., Barbic, M., Blanche, T. J., Bonin, V., Couto, J., Dutta, B., Gratiy, S. L., Gutnisky, D. A., Häusser, M., Karsh, B., Ledochowitsch, P., Lopez, C. M., Mitelut, C., Musa, S., Okun, M., Pachitariu, M., Putzeys, J., Rich, P. D., Rossant, C., Sun, W. L., Svoboda, K., Carandini, M., Harris, K. D., Koch, C., O'Keefe, J., and Harris, T. D. (2017). Fully integrated silicon probes for high-density recording of neural activity. Nature, 551(7679):232 -- 236. [Ko et al., 2013] Ko, H., Cossell, L., Baragli, C., Antolik, J., Clopath, C., Hofer, S. B., and Mrsic-Flogel, T. D. (2013). The emergence of functional microcircuits in visual cortex. Nature, 496(7443):96 -- 100. [Laje and Buonomano, 2013] Laje, R. and Buonomano, D. V. (2013). Robust timing and motor patterns by taming chaos in recurrent neural networks. Nature neuroscience, 16(7):925 -- 33. [Lashley, 1951] Lashley, K. S. (1951). The Problem of Serial Order in Behavior. Cerebral Mechanisms in Behavior, 21:112 -- 146. [Lee et al., 2019] Lee, H., Choi, W., Park, Y., and Paik, S.-B. (2019). Distinct role of flexible and stable encoding in sequential working memory. Neural Networks, 121:419 -- 429. [Leonardo and Fee, 2005] Leonardo, A. and Fee, M. S. (2005). Ensemble Coding of Vocal Control in Birdsong. Journal of Neuroscience, 25(3):652 -- 661. [Litwin-Kumar and Doiron, 2014] Litwin-Kumar, A. and Doiron, B. (2014). Formation and maintenance of neuronal assemblies through synaptic plasticity. Nature Communications, 5(5319):1 -- 12. 20 [Maass, 2016] Maass, W. (2016). Searching for principles of brain computation. Current Opinion in Behavioral Sciences, 11:81 -- 92. [Maass and Markram, 2002] Maass, W. and Markram, H. (2002). Synapses as dynamic memory buffers. Neural Networks, 15(2):155 -- 161. [Mackevicius et al., 2019] Mackevicius, E. L., Bahle, A. H., Williams, A. H., Gu, S., Denisenko, N. I., Goldman, M. S., and Fee, M. S. (2019). Unsupervised discovery of temporal sequences in high-dimensional datasets, with applications to neuroscience. eLife, 8(e38471):1 -- 42. [Manohar et al., 2019] Manohar, S. G., Zokaei, N., Fallon, S. J., Vogels, T. P., and Husain, M. (2019). Neural mechanisms of attending to items in working memory. Neuroscience and Biobehavioral Reviews, 101:1 -- 12. [Mastrogiuseppe and Ostojic, 2018] Mastrogiuseppe, F. and Ostojic, S. (2018). Linking Connectivity, Dynam- ics, and Computations in Low-Rank Recurrent Neural Networks. Neuron, 99(3):609 -- 623.e29. [Morrison et al., 2007] Morrison, A., Aertsen, A., and Diesmann, M. (2007). Spike-timing dependent plasticity in balanced random networks. Neural Computation, 19:1437 -- 1467. [Nicola and Clopath, 2017] Nicola, W. and Clopath, C. (2017). Supervised learning in spiking neural networks with FORCE training. Nature Communications, 8(1):1 -- 15. [Nicola and Clopath, 2019] Nicola, W. and Clopath, C. (2019). A diversity of interneurons and Hebbian plas- ticity facilitate rapid compressible learning in the hippocampus. Nature Neuroscience, 22(7):1168 -- 1181. [Okubo et al., 2015] Okubo, T. S., Mackevicius, E. L., Payne, H. L., Lynch, G. F., and Fee, M. S. (2015). Growth and splitting of neural sequences in songbird vocal development. Nature, 528(7582):352 -- 357. [Park et al., 2017] Park, Y., Choi, W., and Paik, S. B. (2017). Symmetry of learning rate in synaptic plasticity modulates formation of flexible and stable memories. Scientific Reports, 7(1):1 -- 12. [Pastalkova et al., 2008] Pastalkova, E., Itskov, V., Amarasingham, A., and Buzsáki, G. (2008). generated cell assembly sequences in the rat hippocampus. Science, 321(5894):1322 -- 1327. Internally [Peters et al., 2014] Peters, A. J., Chen, S. X., and Komiyama, T. (2014). Emergence of reproducible spa- tiotemporal activity during motor learning. Nature, 510(7504):263 -- 267. [Rajan et al., 2016] Rajan, K., Harvey, C. D., and Tank, D. W. (2016). Recurrent Network Models of Sequence Generation and Memory. Neuron, 90(1):128 -- 142. [Raman et al., 2019] Raman, D. V., Rotondo, A. P., and O'Leary, T. (2019). Fundamental bounds on learning performance in neural circuits. In Proceedings of the National Academy of Sciences. [Rhodes et al., 2004] Rhodes, B. J., Bullock, D., Verwey, W. B., Averbeck, B. B., and Page, M. P. (2004). Learning and production of movement sequences: Behavioral, neurophysiological, and modeling perspectives. Human Movement Science, 23(5):699 -- 746. [Sakai et al., 2003] Sakai, K., Kitaguchi, K., and Hikosaka, O. (2003). Chunking during human visuomotor sequence learning. Experimental Brain Research, 152(2):229 -- 242. [Schaub et al., 2015] Schaub, M. T., Billeh, Y. N., Anastassiou, C. A., Koch, C., and Barahona, M. (2015). Emergence of Slow-Switching Assemblies in Structured Neuronal Networks. PLoS Computational Biology, 11(7):1 -- 28. [Setareh et al., 2018] Setareh, H., Deger, M., and Gerstner, W. (2018). Excitable neuronal assemblies with adaptation as a building block of brain circuits for velocity-controlled signal propagation. PLoS Computational Biology, 14(7):e1006216. [Sjöström et al., 2001] Sjöström, P. J., Turrigiano, G. G., and Nelson, S. B. (2001). Rate, timing, and cooper- ativity jointly determine cortical synaptic plasticity. Neuron, 32(6):1149 -- 1164. [Spreizer et al., 2019] Spreizer, S., Aertsen, A., and Kumar, A. (2019). From space to time: Spatial inhomo- geneities lead to the emergence of spatio-temporal activity sequences in spiking neuronal networks. PLoS Computional Biology, 15(10):e1007432. [Sussillo and Abbott, 2009] Sussillo, D. and Abbott, L. F. (2009). Generating Coherent Patterns of Activity from Chaotic Neural Networks. Neuron, 63(4):544 -- 557. 21 [Tanji, 2001] Tanji, J. (2001). Sequential Organization of Multiple Movements : Involvement of Cortical Motor Areas. Annual Review of Neuroscience, 24:631 -- 651. [Tao, 2013] Tao, T. (2013). Outliers in the spectrum of iid matrices with bounded rank perturbations. Probability Theory and Related Fields, 155(1-2):231 -- 263. [Tully et al., 2016] Tully, P. J., Lindén, H., Hennig, M. H., and Lansner, A. (2016). Spike-Based Bayesian- Hebbian Learning of Temporal Sequences. PLoS Computational Biology, 12(5). [Vogels et al., 2011] Vogels, T. P., Sprekeler, H., Zenke, F., Clopath, C., and Gerstner, W. (2011). Inhibitory Plasticity Balances Excitation and Inhibition in Sensory Pathways and Memory Networks. Science, 334(De- cember 2011):1569 -- 1573. [Waddington et al., 2012] Waddington, A., de Kamps, M., and Cohen, N. (2012). Triphasic spike-timing- dependent plasticity organizes networks to produce robust sequences of neural activity. Frontiers in Compu- tational Neuroscience, 6:1 -- 14. [Werbos, 1990] Werbos, P. J. (1990). Backpropagation Through Time: What It Does and How to Do It. Proceedings of the IEEE, 78(10):1550 -- 1560. [Zenke et al., 2015] Zenke, F., Agnes, E. J., and Gerstner, W. (2015). Diverse synaptic plasticity mechanisms orchestrated to form and retrieve memories in spiking neural networks. Nature Communications, 6(6922):1 -- 13. [Zheng and Triesch, 2014] Zheng, P. and Triesch, J. (2014). Robust development of synfire chains from multiple plasticity mechanisms. Frontiers in Computational Neuroscience, 8:1 -- 10. 22 Supplementary material Fig S1. Randomly switching dynamics. The recurrent network is stimulated with external input that is spatially clustered, but temporally uncorrelated. Each cluster is stimulated for 50 ms, with 50 ms gaps in between stimulations. The rate of external stimulation is 18 kHz during training. This is repeated for 20 minutes after which the network stabilizes during 20 minutes of spontaneous activity. (A) A diagonal structure is learned in the recurrent weight matrix. Since there are no temporal correlations in the external input, there is no off-diagonal structure. (B) The spectrum shows an eigenvalue gap. This indicates the emergence of a slower time scale. The leading eigenvalues do not have an imaginary part, pointing at the absence of feedforward structure and thus there is no sequential dynamics. (C) Under a regime of spontaneous dynamics (i.e. uncorrelated Poisson inputs), the clusters are randomly reactivated. 23 ADetail of recurrent weight matrix(5 first clusters) 8016024032080160240320Postsynaptic neuron indexPresynaptic neuron indexBC-5000500Real part-1000-50005001000Imaginary partSpectrum of RNN weight matrix(full matrix)800160024003000Recurrent networkneuron index Random switching between clusters during spontaneous dynamicsTime [s]0.210.6 Fig S2. The connectivity structure is stable under spontaneous dynamics. (A) After 60 minutes of training, the network stabilizes during spontaneous activity. During the first 30 minutes of spontaneous dynamics, the connectivity still changes. More specifically, the imaginary parts of the leading eigenvalues increase. This leads to a higher switching frequency and as such a smaller period in the sequential activity. After around 30 minutes, a fixed point is reached. The first row shows spike trains at different times, for one second of spontaneous activity. The second row shows the spectra of the weight matrix at those times. (B) After 60 minutes of sequential stimulation, we test reinforcement and degradation of the learned connectivity by decoupling the plasticity from the dynamics. We plot the evolution of the off-diagonal weights during spontaneous dynamics in two separate cases: (i) we run the dynamics of the network using a frozen copy of the learned weight matrix and apply plastic changes that result from the dynamics to the original weight matrix (blue curve); (ii) we run the dynamics of the network using a frozen copy of the learned weight matrix where the off-diagonal structure was removed and apply plastic changes that result from the dynamics to the original weight matrix (red curve). We can see that in the former, the learned connectivity is reinforced and in the latter, the learned connectivity degrades. Off-diagonal weights (the y-axis) are quantified by averaging over the weights in the 80 by 80 blocks in the lower diagonal, for the 30 different clusters. The curves are the means over the 30 clusters and the error bars one standard deviation. 24 Imaginary axis 10002000-2000-100000800160024003000Recurrent networkneuron index 00.51Time [s] 800160024003000800160024003000Development and stability of the temporal backbone0204060Time [min]1.522.5Off-diagonalweightsDegradation and reinforcement of connectivity during spontaneous dynamicsReal axis -5000500After 60 minutes ofspontaneous dynamics Real axis -5000500Real axis -5000500After 30 minutes ofspontaneous dynamics After 60 minutes ofsequential stimulation 00.51Time [s] 00.51Time [s] SpectrumSpectrumSpectrumEnd of trainingSpontaneous random activityABDegradationReinforcement10002000-2000-10000010002000-2000-100000 Fig S3. Noisy learning. The sequence ABCBA is relearned four times for 12 seconds each. Before relearning, the read-out weight matrix W RE was always reset. When active, read-out neurons fire two spikes on average +/− one spike. This variability is a consequence of the noisy learning process. Fig S4. Details of some network properties. (A) Duration that a cluster is activated as a function of network size (B) Raster plot of sequential dynamics for NE = 1200 and NC = 40, after training. We observe that by reducing the cluster size, the irregularities in the sequential dynamics are increased (compare with Fig 2). (C) Two raster plots showing two different levels of robustness (summary plot in Fig 5B). In both cases, at t = 1s (purple arrow), 40 read-out synapses are deleted for each cluster. Left panel: NC = 120, each read-out neuron fires two spikes before deletion and one spike after deletion resulting in ∼ 50% performance. Right panel: NC = 200, each read-out neuron fires two spikes before deletion and one or two spikes after deletion resulting in a higher performance (∼ 80%). 25 00.20.40.60.8Time [s]ABCRead-out neuron00.20.40.60.8Time [s]ABCRead-out neuron00.20.40.60.8Time [s]ABCRead-out neuron00.20.40.60.8Time [s]ABCRead-out neuron012200040006000Recurrent networkneuron index012Time [s]ABCRead-out neuronSyanp Duration of cluster activations in various networks Networks with small clusters show more irregular sequential dynamicsC Spike rasters showing in more detail the robustness experiment for two casesABNC = 120NC = 200Synapse deletionSynapse deletion Fig S5. The role of inhibition. (A) Inhibitory neurons are necessary to prevent pathological excitatory activity. (B) The weights projecting from the inhibitory neurons to the excitatory neurons without inhibitory plasticity are random (left panel). The weights projecting from the inhibitory neurons to the excitatory neurons with inhibitory plasticity shows some structure (right panel). (C) The full spectrum of the recurrent weight matrix after learning without inhibitory plasticity. (D) Without inhibitory plasticity, the sequential dynamics shows irregularities. The inhibitory plasticity allows for better parameters to be found to stabilize the sequential dynamics in the recurrent network. Fig S6. Sensitivity to parameters. The periods of the sequential dynamics are computed after one hour of external stimulation and one hour of spontaneous dynamics. Only one parameter at a time is changed. (A) The adaptation time constant is varied. (B) The time gap between external sequential stimulations is varied. (C) The time constants of the voltage-based STDP rule are varied. The lines are guides to the eye and the error bars indicate one standard deviation. 26 04080Time [ms]180016002400Recurrent networkneuron indexA No inhibitory neurons leads to pathological activity-5000500Real axis-2000-1000010002000Imaginary axis00.511.52Time [s]800160024003000Recurrent networkneuron indexC Spectrum after learning without inhibitory plasticityD Spike raster shows irregularities15080Presynaptic Ineuron index180160240320400Postsynaptic Eneuron index0102030405060150100Presynaptic Ineuron index180160240320400Postsynaptic Eneuron index050100150200B Inhibitory to excitatory weights without and with inhibitory plasticity 357Pairing gap [ms]440460480Period [ms]-0.50+0.5-10+1-20+2Change in time constant [ms]420440460480500Period [ms]vxu50100150Adaptation timeconstant [ms]440460480Period [ms]CAB
1301.7747
1
1301
2013-01-31T20:47:53
Entrainment of the intrinsic dynamics of single isolated neurons by natural-like input
[ "q-bio.NC" ]
Neuronal dynamics is intrinsically unstable, producing activity fluctuations that are essentially scale-free. Here we show that while these scale-free fluctuations are independent of temporal input statistics, they can be entrained by input variation. Joint input output statistics and spike train reproducibility in synaptically isolated cortical neurons were measured in response to various input regimes over extended time scales (many minutes). Response entrainment was found to be maximal when the input itself possesses natural-like, scale-free statistics. We conclude that preference for natural stimuli, often observed at the system level, exists already at the elementary, single neuron level.
q-bio.NC
q-bio
Entrainment of the intrinsic dynamics of single isolated neurons by natural-like input Asaf Gal1,2,∗, Shimon Marom1 1 Network Biology Research Laboratories, Lorry Lokey Interdisciplinary Center for Life Sciences and Engineering, Technion, Haifa, Israel 2 The Interdisciplinary Center for Neural Computation (ICNC), The Hebrew University, Jerusalem, Israel ∗ E-mail: [email protected] Abstract Neuronal dynamics is intrinsically unstable, producing activity fluctuations that are essentially scale-free. Here we show that while these scale-free fluctuations are independent of temporal input statistics, they can be entrained by input variation. Joint input output statistics and spike train reproducibility in synaptically isolated cortical neurons were measured in response to various input regimes over extended time scales (many minutes). Response entrainment was found to be maximal when the input itself possesses natural-like, scale-free statistics. We conclude that preference for natural stimuli, often observed at the system level, exists already at the elementary, single neuron level. Introduction Variability is a most prominent property of neural activity and neural response: neurons and neural networks behave in an irregular and indeterministic manner both spontaneously, and in response to series of stimuli. At the single neuron level, variability is observable in practically all aspects of evoked activity: irregularity of the spike train, trial-to-trial variability in spike counts, as well as irreproducibility of train structure evoked by identical input series (Faisal et al., 2008; Yarom and Hounsgaard, 2011). Generally, variability in responses to repeated presentation of a 1 stimulus is a significant constraint on information carrying and processing capacity. However, as demonstrated in several cases, response variability might be quenched by a variation introduced to the input itself (Churchland et al., 2010). At the single neuron level, it was demonstrated that when stimulated with constant input, the neuronal spike train differ substantially from trial to trial (Bryant and Segundo, 1976; Mainen and Sejnowski, 1995). In contrast, when stimulated with a fluctuating (filtered white noise) input, the reproducibility of the spike train is dramatically improved to the point of perfect repeatability, locking itself to (i.e. entrained by) input fluctuations, reliably encoding its structure. This key property was reproduced in a stochastic simulation of a Hodgkin-Huxley neuron, relating it to the properties of the underlying ion channels (Schneidman et al., 1998). Other works have demonstrated the existence of repeatable spike patterns under different types of stimuli (Fellous et al., 2004). While these measurements and simulations were limited to timescale of seconds, it is known that when neuronal activity is observed over extended time scales, slower effects are recruited and excitability dynamics becomes rich (Marom, 2010). In a recent work (Gal et al., 2010) we have shown that, indeed, when presented with long (> 1h) sequences of pulse stimuli, single neuron response dynamics becomes intermittent and irregular, exhibiting scale-free fluctuations (e.g. with auto-correlation that lacks a characteristic scale) and transitions between quasi-stable response pattern modes. Given these slower modulatory processes, it is not obvious that the statistically unstructured random input series, that are capable of quenching response variation over limited time scales, will effectively entrain response variability over extended durations (minutes and more). The biophysical mechanism underlying the capacity of unstructured random input to entrain response variability relies on matching between time scales of input variations and time scales of the stochastic processes that generate the action potential. In contrast, when longer stochastic processes are allowed, they are left unmatched by the above unstructured input. It is therefore natural to hypothesize that in order to entrain neuronal response over extended time scales, the variations of the input series must match the scale-free temporal structure of intrinsic neuronal response dynamics. Indeed, at least 2 at the system level, neuronal response variability is reduced under natural or natural-like sensory input (Aertsen and Johannesma, 1981; Baddeley et al., 1997; de Ruyter van Steveninck et al., 1997; Garcia-Lazaro et al., 2006, 2011; Yu et al., 2005). These natural-like signals are often characterized by long range temporal correlations, and a general scale-free temporal structure (De Coensel et al., 2003; Simoncelli, 2003; Voss and Clarke, 1975). In this study we directly measure the impacts of input temporal structure on response variability over extended time scales in isolated cultured cortical neurons. We show that while the response of neurons is temporally scale-free, independently of input statistics, entrainment is maximal when the input itself has a matching, scale-free structure. We also perform analogous analyses to those of Mainen and Sejnowski, quantifying the reproducibility of spike trains under different types of input. Here too, natural-like input minimizes the trial-to-trial variability of the spike train. We conclude that the rich and complex neuronal dynamics enable the neuron to match its dynamics to that of the natural environment, and that "tuning" to natural input statistics arises already at the atomic level of neural processing. Materials and Methods Culture preparation. Cortical tissues were obtained from newborn (< 24h) rats (Sprague- Dawley) and dissociated following procedures described in earlier studies (Marom and Shahaf, 2002). The cells were plated directly onto substrate-integrated multi-electrode arrays (MEA) and developed for a time period of 2-3 weeks before their use. A total amount of approxi- mately 106 cells was seeded on poly-ethyene-immine (PEI) coated MEAs. The preparations were kept in Minimal Essential Medium (MEM) supplemented with heat-inactivated horse serum (5%), glutamine (0.5mM ), glucose (20mM ), and gentamycin (10µg/ml), and main- tained in an atmosphere of 37◦C, 5% CO2 and 95% air in an incubator as well as during the recording phases. An array of Ti/Au/TiN extracellular electrodes, 30µm in diameter, and spaced either 500µm or 200µm from each other (MultiChannelSystems - MCS, Reutlin- gen, Germany) were used. Synaptic transmission in the network was completely blocked by 3 adding 20µM APV (amino-5-phosphonovaleric acid), 10µM CNQX (6-cyano-7-nitroquinoxaline- 2,3-dione), and 5µM Bicuculline-methiodide to the bathing solution. All experiments were performed in accordance with the regulations (and under the supervision) of the Technion - Israel Institute of Technology animal care committee. Measurements and stimulation. A commercial amplifier (MEA-1060-inv-BC, MCS) with frequency limits of 150-3000Hz and a gain of x1024 was used. Rectangular 200µs biphasic 600- 800mV voltage stimulation through extracellular electrodes was performed using a dedicated stimulus generator (STG4004, MCS). In the context of this study, no difference was observed in the behavior of neurons under current or voltage stimulation. Data was digitized to 16bit using a USB-ME256 system (MCS). Each recorded channel was sampled at a frequency of 20KHz. One hour after the addition of synaptic blockers, the stimulation electrode was selected as one evoking well-isolated spikes with high signal to noise ratio, in as many recording electrodes as possible. From the selected recording electrodes, voltage traces of 15-20ms post stimulus were collected. Spike detection was performed off-line by a manual threshold-based procedure. A 3ms long spike shape was extracted for each response for further noise cleaning and analysis. Stability of spike shape and activity dynamics criteria were applied in order to validate experimental stability, as described in (Gal et al., 2010). Random stimulation sequences were generated by modulating a constant stimulus interval sequence with a noise signal. This noise signal was either a white Gaussian noise or 1/f Gaussian noise generated by weighting the frequency components of the Gaussian white noise. In both cases a low cutoff was applied to have a minimum interval of 20ms. The SD of the noisy stimulation interval sequence was set such that its CV will match the CV of its response constant interval. For example, if a neuron responded to a 100ms constant interval sequence with an average inter spike interval of 200ms and SD of 40ms, the SD of the noisy interval sequence was set to 20ms. Data analysis. Analysis throughout this study was performed on either the spike time series, or on a smoothened firing rate time series, produced by filtering the spike train with a sliding 4 rectangular window (Figure 1C). Spectral analysis was performed on a binned time series, with a bin size of 1s. The Power Spectrum Density (PSD) was estimated using a modified Periodogram. The Allan variance, which is commonly used to identify fractal point processes (Lowen and Teich, 1996, 2005) was calculated by binning the time series with different bin sizes T . For each binned time series, the Allan variance is defined as A(T ) = <ZT (k)−ZT (k+1)>2 , where ZT is the binned 2·<ZT > series. Another measure, widely used to characterize the temporal statistics of long memory and non stationary time series, is Detrended Fluctuation Analysis (DFA, Peng et al., 1995), in which the fluctuations (in terms of mean square error) around a piecewise linear fit to the time series are quantified, for different segment durations. Since both the Periodogram and the Allan variance are can be regarded as power law for the purpose of this work, power law functions were fitted to the tail of these curves, and their exponents were used as descriptive statistics (see Gal et al., 2010). The similarity between two spike trains was assessed with two distance measures: (1) A correlation based metric, defined as one minus the correlation between binned time series (Za,Zb): d = 1 − corr(Za, Zb). This is a rate-based measure, which is insensitive to temporal features of the spike train below resolution dictated by the bin size. (2) The Victor-Purpura spike train metric (Toups et al., 2011; Victor, 2005; Victor and Purpura, 1997), which defines the distance between the two spike trains as the minimal cost of transforming one into the other. Briefly, a spike train is modified by a combination three possible steps: inserting or deleting a spike, with a cost of +1, and moving a spike in time, with a cost of q·dt , where q is a temporal resolution parameter. The value of q sets the sensitivity of the metric to fine temporal features, and is set here to a default value of q = 1s. The dependence of the results on the value of q is shown in Figure 4. The Victor-Purpura metric was calculated using Matlab code downloaded from the website of Jonathan Victor. 5 Results Unexplained response variability is minimized by scale-free input In this paper we investigate the response properties of individual neurons, independent of synap- tic and network effects. To that end, experiments are conducted on cultured cortical neurons, functionally isolated from their network by means of pharmacological block of both glutamatergic and GABAergic synapses (see Gal et al., 2010, and Materials and Methods). Individual neurons are stimulated with sequences of short, identical extracellular electrical pulses. In response to a single pulse, a neuron either responds by emitting a spike, or fails to do so (Figure 1A); neuronal responses are monitored by extracellular recording electrodes. As previously reported (Gal et al., 2010), when repeatedly stimulated over extended durations (minutes and more) with a low (∼1Hz) stimulation rate, a neuron responds to each and every stimulation pulse (1:1 mode). As stimulation rate increases, the excitability of the neuron declines, and at a certain, neuron- specific critical stimulation rate, the 1:1 response mode breaks, and the neuron exhibits rich and complex response dynamics (Figure 1B). Here, we study the properties of response dynamics evoked by three statistically different regimes of stimulation series: (i) constant interval regime, (ii) white noise regime in which the interval series is modulated by a Gaussian white noise process, and (iii) scale-free regime in which the intervals are modulated by a 1/f β process, with parameter β = 1, which is more representative of natural sensory input (De Coensel et al., 2003; Simoncelli, 2003; Voss and Clarke, 1975) and similar to activity properties of cortical neurons in-vivo (Bhattacharya et al., 2005; Lowen et al., 2001; Lowen and Teich, 1993, 1996; Teich et al., 1997). The term scale-free is used here to designate a timeseries with an autocorrelation that decays slowly, usually as a power-law without a typical scale. White noise on the other hand, which has a delta-function autocorrelation, is regarded as a zero scale signal. The choice of β = 1 is typical to a wide range of natural signals, in natural environments and in biological systems. As will be shown, the results of this work are not sensitive to the exact value of β, consistent with previous studies (Garcia-Lazaro et al., 2006). Interval sequences in all stimulation regimes were normalized to 6 have the same mean and standard deviation (the latter is applicable only to the second and third regime). The mean stimulation rate (the reciprocal of the mean stimulation interval) was set to a high enough value to drive the neuron beyond its critical point, leading to response failures and intermittency (Gal et al., 2010). The interval standard deviation (SD) was chosen to approximately match the intrinsic SD of the response to constant interval stimulation. Figure 1D shows examples of extracts from the three stimulation regimes, as well as their autocorrelation functions and the Power Spectral Densities (PSD). Our analysis starts by stimulating neurons with long sequences (> 1h) of each of the stim- ulation regimes. Figure 2A shows an example of the response of the same neuron to the three regimes. It is immediately obvious that the three input regimes did not cause a significant dif- ference in the statistical properties of the responses. This is formally shown in the plots of the PSD, Allan variance, detrended Fluctuation Analysis (DFA), rate histograms and inter-spike interval (ISI) histograms (Figure 2B-F). Clearly, the macroscopic properties of the evoked spike trains are the same under all stimulation regimes: the neuron exhibits scale-free dynamics, char- acterized by power law statistics, in accordance with previously published analysis (Gal et al., 2010). In order to confirm the insignificance of the differences between the different statistical measures under the three regimes, a control experiment was performed, in which 4 neurons were stimulated and recorded, and each stimulation block was repeated 10 times. For each statistical analysis presented in Figure 2B-F, the relevant parameter was estimated independently from the response to each repetition. The value ranges of these parameters from a single neuron are depicted in the insets of each panel, and show that indeed the temporal statistics are the same under the three regimes. We emphasize that the claim made here is not that the specific model chosen for each statistics is the correct one (i.e. that the PSD is an exact power law), rather that these fits are good enough representative shapes, useful for comparing the population statistics. The above results imply that the various membrane and cellular processes underlying the stochastic response fluctuations are effectively insensitive to the statistical structure of the input regime. 7 1/f -type response statistics can be interpreted as a modulation of neuronal excitability by a cascade of oscillating processes at various time scales. An oscillator can be entrained (i.e. phase- locked) by a driving stimulus at a frequency that matches the oscillator's natural frequency, and with a magnitude proportional to the oscillation amplitude. This suggests that the 1/f intrinsic fluctuations could be entrainable by a matching 1/f stimulation. Figure 3 demonstrates this effect: The stimulation and response of a neuron are plotted in two timescales, for the three stimulation regimes. On the short timescale (panel A, 5s bin size), the response in the white noise and scale-free regimes nicely follows the stimulation, while in the constant interval regime, the input has no bearing on timing of the output fluctua- tions. On a longer timescale (panel B, bin size of 100s), the white noise input is practically flattened, becoming constant; as a result, it fails to entrain the response. In contrast, in the scale-free regime entrainment is evident throughout. Figure 3C shows a scatter plot of the input and output rates of the neuron, calculated with a 5s bin size. As expected from the data of Figure 2, the input and output ranges of the white-noise (red) and scale-free (green) are practically identical. However, the correlation in the scale-free data is much higher, as expressed in the reduced variability around the linear trend line. This indicates that indeed the scale-free fluctuations in neuronal dynamics are best entrained by scale-free input. For 13 of the 17 recorded neurons, the input output correlation coefficient significantly increases under scale-free input regime, compared to white noise regime. A Wilcoxon signed-rank test yields p < 0.01 for the effect of the input regime on input/output correlation. Thus, under scale-free input, there is no change in the amount of variability in neuronal response, but more of it is explained by the input. Figure 3D shows a scaled correlation analysis: The correlation between input and output is calculated for responses in white noise and scale-free regimes on different timescales. The correlation in a given timescale T is calculated by smoothing the stimulation and response series with a rectangular window of width T , and subtracting from it a version of the series, smoothed with a window of size 25T . This effectively band-passes the time series around T . While the correlation in the short timescale regime is similar for the white noise and 8 scale free regimes (the apparent increase for white noise is non typical, a Wilcoxon signed-rank test yields insignificant difference), for longer timescales it decreases in the case of white noise compared to scale free (p < 0.01, Wilcoxon signed-rank test). The above results show that there is a significant increase in correlation upon change from white noise (i.e. β = 0) to scale free (β = 1) input. However, there is nothing unique about β = 1: other types of inputs might serve just as well. In order to asses the sensitivity of the increased input-output correlation to the value of β, neurons (n=31) were exposed to 9 stim- ulation blocks, each characterized by a different value of β, ranging from 0 to 2, and lasting 70 minutes. For each block, the correlation between input and response was computed, and compared with the correlation for β = 0. The population statistics of this correlation ratio are depicted in Figure 3E, and show a clear concave shape, with a peak in the mid-range values, and a decrease toward lower and higher values of β. A similar peak is observed when plotting the distribution of preferred exponents (i.e. the exponents which results in the strongest correlation for each neuron, Figure 3F). These peaks around β = 1, however, are wide, suggesting that the variability between neurons is considerable, and that inputs characterized by a wide range of exponents are equally effective in entraining neuronal responses. The decrease of the correlation for input with large exponents, which are dominated by slow oscillations, is important for the understanding of this phenomena. Figure 3G shows extracts from the response to 1/f stimulation (left) and 1/f 2 stimulation (right). It is easy to see how, for 1/f 2, the faster fluctuations remain un-entrained, while slower oscillations are nicely locked by modulations in the input. Neuronal response repeatability is maximized by scale-free input Another functionally relevant aspect of the entrainment described above, concerns response repeatability. It is well established that in spite of the extensive variability of neuronal responses, spike train structure is repeatable when the input is fluctuating, both at the single neuron and sensory system levels (Churchland et al., 2010; de Ruyter van Steveninck et al., 1997; Mainen 9 and Sejnowski, 1995). In the following set of experiments we ask whether a scale-free stimulation regime enhances repeatability in general, and over long time scales in particular. To this aim we stimulated neurons with 10 repetitions of the same input sequence under each of the three stimulation regimes: constant intervals, white noise and scale-free. Each sequence lasted 10 minutes, separated by a 10 minute break. Figure 4A shows responses of one neuron to ten identical sequences under each stimulation regime. It is immediately obvious that although some reproducibility exists under the white noise regime, the reproducibility of responses to scale-free sequences is much higher. This is a direct consequence of the previous section analysis: if responses are more correlated to the input, they will be more correlated between themselves under repetitions of the same input sequence. This observation is quantified using two kinds of spike train similarity measure: a rate-based measure (the correlation distance between spike histograms, calculated with 1s bins; see Methods section), and a time-based measure (the Victor- Purpura distance, Toups et al., 2011; Victor, 2005; Victor and Purpura, 1997). Figure 4B depicts the pairwise distance matrices of the responses of Figure 4A, calculated using these two metrics. Responses to white noise stimulation are significantly more reproducible than responses to constant input; this is in agreement with (Mainen and Sejnowski, 1995). But as expected from the results presented above (Figure 3) the responses to the scale-free sequences are significantly more reproducible than those of white noise input. The purple lines of Figures 4C and 4D depict the mean and SD of the values in the distance matrices shown in Figure 4B. Of 14 recorded neurons, 11 neurons showed significant improvement in reproducibility for scale-free sequences compared to constant interval and white noise, as quantified by the two metrics (there were no cases of disagreement). The grey lines depict the trends of mean distances calculated for all of the 14 recorded neurons. A Wilcoxon signed-rank test for the mean pair-wise distances shows an overall significant effect for the input regime on response repeatability (p < 0.01 for both white noise vs. constant and for scale-free vs. white noise, for both metrics). The results are insensitive to the choice of the temporal parameter q of the spike train metric (Figure 4F), which may point to the lack of characteristic scale in this phenomenon. 10 The above enhanced reproducibility does not stem from a decreasing spike rate. Next to the preservation of the mean stimulation rate in all sequences, the spike rate itself does not decrease for scale-free input (see Figure 4E). For all of the 14 cells recorded, there was no significant decrease in firing rate under the scale-free regime compared with constant and white noise. Since the scale-free input itself is structured, it is expected that even a "Bernoulli" neuron that has a constant probability of response to a pulse stimulation, will have some reproducibility of its output spike train. Figure 5A shows a comparison between metric analyses on actual responses (grey), and on a surrogate Bernoulli neuron (purple) that responds with a constant probability (set to the mean response probability of the real neuron). As expected, the responses of the simulated neuron are indeed more reproducible for scale-free input, but not as consistent as the real neuron. It is also possible to construct a more detailed neuronal response curve, which takes into account the dependency of the response probability on the last interval between stimuli (an example for such a curve is given in Figure 5B). Interestingly, the curves of the white noise and scale-free regimes substantially differ, pointing to a strong history dependence in response probability. As shown in Figure 5A, the resulting metric analyses (orange) behave more like the real neuron when compared to the Bernoulli neuron, yet cannot account for the entire effect. It is reasonable to believe that one might construct a response model that takes into account deeper history of the stimulation and response sequences, to produce better fitting. It should be emphasized though, that while these neuronal response models do reproduce the repeatability effect to a significant extent, they do not reproduce the intrinsic scale-free fluctuations, and can not be considered as successful explanatory models. A scale-free input is characterized by the abundance of relatively long "breaks" in stimulation, or long periods with low stimulation rate, enabling recovery of internal processes from previous activations. As repeatedly shown over the past 15 years, the longer a neuron is exposed to repeated activations, longer recovery times are required (Ellerkmann et al., 2001; Fairhall et al., 2001; Lundstrom et al., 2008; Marom, 2009; Toib et al., 1998). A scale-free input statistics is inherently matched to such a mechanistic context: elongating a scale-free input series naturally 11 gives rise to longer breaks, hence allowing for stabilization on every scale. An illustration of this property is provided by reanalyzing the data of Figure 4 over blocks of increasing lengths. The results are summarized in Figures 5C and 5D, showing that the accumulation of variability (or divergence of response) with increasing block size (1-10 minutes range) is significantly slower than its accumulation in responses to other stimulation regimes. Discussion In this paper we have shown how a natural-like, scale-free input entrains fluctuations of single neuron responses over extended timescales. We have demonstrated this property by comparing neuronal responses in three different stimulation regimes: constant interval, white noise and scale-free. In the case of the scale-free regime, the correlation between the input and the re- sponse is significantly higher, and the repeatability of response is considerably enhanced. These characteristics are stable over long, practically unlimited durations. While the results do show a preference to mid-range values of β (around 1), there is nothing special about the exact value; what seems to be important is that the entrainment decreases when the slow frequency compo- nent in the input becomes either too dominant (large β) or marginal (low β). It has long been acknowledged that responsiveness of neural systems is optimized to the ranges of statistics found in natural inputs (Aertsen and Johannesma, 1981; Baddeley et al., 1997; de Ruyter van Steveninck et al., 1997; Garcia-Lazaro et al., 2006, 2011; Yu et al., 2005). Here we show that preference to natural statistics is not limited to large-scale neural systems; rather, it goes all the way down, to the atomic level of neural organization, namely the single, isolated neuron. At the shorter timescales, the entrainment of neuronal fluctuations by white noise input is explained by the fast stochastic processes underlying the generation of action potentials (Schnei- dman et al., 1998). There, a variation in input allows for recovery of inactivation processes, unlike the case of a constant input that drives the neuron to operate around the limit of channel availability threshold, making it highly sensitive to stochastic events. It is reasonable to assume 12 that a similar explanation would also be appropriate over extended timescales: a long break (or a period of low-rate stimulation) enables recovery of slow processes, in contrast to constant input (or shortly correlated input) that drives these processes to a highly stochastic operation point. It is natural to assume that such processes include slow inactivation properties of the ionic channels themselves (Ellerkmann et al., 2001; Marom, 2009; Soudry and Meir, 2010; Toib et al., 1998), or other cellular modulatory processes (e.g., protein phosphorylation, protein synthesis and metabolic cycles). From the more abstract, functional point of view, when the temporal structure of the input is relatively dull, it can only entrain a narrow range of cellular processes underlying neuronal dynamics. Under these conditions, a large fraction of the response variability is tagged "un- explained". However, when the input is temporally rich, it matches the temporal manifold complex structure of the intrinsic dynamics, and the former "unexplained" variability becomes information-carrying. Thus, a neuron can be viewed as a collection of entangled information channels, distributed over a continuum of timescales. In this picture, information transfer is maximized when there is information to be transferred on any given scale. Acknowledgments The authors thank Danni Dagan, Danny Eytan and Avner Wallach for helpful discussions and insightful input, and Eleonora Lyakhov and Vladimir Lyakhov for invaluable technical assistance in conducting the experiments. The research leading to these results has received funding from the European Unions Seventh Framework Program FP7 under grant agreement 269459 and was also supported by a grant of the Ministry of Science and Technology of the State of Israel and MATERA grant agreement 3-7878. References Aertsen A, Johannesma P (1981) A comparison of the spectro-temporal sensitivity of auditory neurons to tonal and natural stimuli. Biological Cybernetics 42:145 -- 156. 13 Baddeley R, Abbott LF, Booth MC, Sengpiel F, Freeman T, Wakeman EA, Rolls ET (1997) Responses of neurons in primary and inferior temporal visual cortices to natural scenes. Proc. R. Soc. Lond. B 264:1775 -- 83. Bhattacharya J, Edwards J, Mamelak aN, Schuman EM (2005) Long-range temporal correlations in the spontaneous spiking of neurons in the hippocampal-amygdala complex of humans. Neuroscience 131:547 -- 55. Bryant H, Segundo J (1976) Spike initiation by transmembrane current: a white-noise analysis. The Journal of Physiology pp. 279 -- 314. Churchland MM, Yu BM, Cunningham JP, Sugrue LP, Cohen MR, Corrado GS, Newsome WT, Clark AM, Hosseini P, Scott BB, Bradley DC, Smith Ma, Kohn A, Movshon JA, Armstrong KM, Moore T, Chang SW, Snyder LH, Lisberger SG, Priebe NJ, Finn IM, Ferster D, Ryu SI, Santhanam G, Sahani M, Shenoy KV (2010) Stimulus onset quenches neural variability: a widespread cortical phenomenon. Nature Neuroscience 13:369 -- 78. De Coensel B, Botteldooren D, De Muer T (2003) 1/f Noise in Rural and Urban Soundscapes. Acta Acustica united with Acustica 89:9. De Ruyter Van Steveninck RR, Lewen GD, Strong SP, Koberle R, Bialek W (1997) Repro- ducibility and variability in neural spike trains. Science 275:1805 -- 1808. Ellerkmann RK, Riazanski V, Elger CE, Urban BW, Beck H (2001) Slow recovery from inacti- vation regulates the availability of voltage-dependent Na(+) channels in hippocampal granule cells, hilar neurons and basket cells. The Journal of Physiology 532:385 -- 97. Fairhall AL, Lewen GD, Bialek W, de Ruyter Van Steveninck RR (2001) Efficiency and ambi- guity in an adaptive neural code. Nature 412:787 -- 792. Faisal AA, Selen LPJ, Wolpert DM (2008) Noise in the nervous system. Nature Reviews Neuroscience 9:292 -- 303. 14 Fellous J, Tiesinga P, Thomas P, Sejnowski T (2004) Discovering spike patterns in neuronal responses. The Journal of Neuroscience 24:2989 -- 3001. Gal A, Eytan D, Wallach A, Sandler M, Schiller J, Marom S (2010) Dynamics of excitability over extended timescales in cultured cortical neurons. The Journal of Neuroscience 30:16332 -- 42. Garcia-Lazaro JA, Ahmed B, Schnupp JWH (2006) Tuning to natural stimulus dynamics in primary auditory cortex. Current Biology 16:264 -- 271. Garcia-Lazaro Ja, Ahmed B, Schnupp JWH (2011) Emergence of Tuning to Natural Stimulus Statistics along the Central Auditory Pathway. PLoS ONE 6:e22584. Lowen SB, Ozaki T, Kaplan E, Saleh BE, Teich MC (2001) Fractal features of dark, maintained, and driven neural discharges in the cat visual system. Methods 24:377 -- 394. Lowen SB, Teich MC (1993) Fractal renewal processes generate 1/f noise. Physical Review E 47:992 -- 1001. Lowen SB, Teich MC (1996) The periodogram and Allan variance reveal fractal exponents greater than unity in auditory-nerve spike trains. The Journal of the Acoustical Society of America 99:3585 -- 3591. Lowen SB, Teich MC (2005) Fractal-based point processes Wiley-Interscience, Hoboken, N.J. Lundstrom BN, Higgs MH, Spain WJ, Fairhall AL (2008) Fractional differentiation by neocor- tical pyramidal neurons. Nature Neuroscience 11:1335 -- 1342. Mainen ZF, Sejnowski TJ (1995) Reliability of spike timing in neocortical neurons. Sci- ence 268:1503 -- 1506. Marom S (2009) Adaptive transition rates in excitable membranes. Frontiers in Computational Neuroscience 3:2. Marom S (2010) Neural timescales or lack thereof. Progress in Neurobiology 90:16 -- 28. 15 Marom S, Shahaf G (2002) Development, learning and memory in large random networks of cortical neurons: lessons beyond anatomy. Quarterly Reviews of Biophysics 35:63 -- 87. Peng CK, Havlin S, Stanley HE, Goldberger AL (1995) Quantification of scaling exponents and crossover phenomena in nonstationary heartbeat time series. Chaos 5:82. Schneidman E, Freedman B, Segev I (1998) Ion channel stochasticity may be critical in deter- mining the reliability and precision of spike timing. Neural Computation 10:1679 -- 1703. Simoncelli EP (2003) Vision and the statistics of the visual environment. Current Opinion in Neurobiology 13:144 -- 149. Soudry D, Meir R (2010) History-dependent Dynamics in a Generic Model of Ion Channels - an Analytic Study. Frontiers in Computational Neuroscience 4:1 -- 19. Teich MC, Heneghan C, Lowen SB, Ozaki T, Kaplan E (1997) Fractal character of the neu- ral spike train in the visual system of the cat. Journal of the Optical Society of America A 14:529 -- 546. Toib A, Lyakhov V, Marom S (1998) Interaction between duration of activity and time course of recovery from slow inactivation in mammalian brain Na+ channels. The Journal of Neu- roscience 18:1893 -- 1903. Toups JV, Fellous JM, Thomas PJ, Sejnowski TJ, Tiesinga PH (2011) Finding the event struc- ture of neuronal spike trains. Neural Computation 23:2169 -- 208. Victor JD (2005) Spike train metrics. Current opinion in neurobiology 15:585 -- 92. Victor JD, Purpura KP (1997) Metric-space analysis of spike trains: theory, algorithms and application. Network: Computation in Neural Systems 8:127 -- 164. Voss RF, Clarke J (1975) 1/f noise in music and speech. Nature 258:317 -- 318. Yarom Y, Hounsgaard J (2011) Voltage fluctuations in neurons: signal or noise? Physiological Reviews 91:917 -- 29. 16 Yu Y, Romero R, Lee T (2005) Preference of Sensory Neural Coding for 1/f Signals. Physical Review Letters 94:18 -- 21. 17 Figure captions Figure 1. Data analysis and stimulation regimes. (A) Examples of voltage traces recorded following several stimulation pulses, delivered at 15Hz. In response to such a pulse, a neuron sometimes emits a spike (blue traces) or fail to do so (red traces). Stimulation pulses are 400µs wide and start at t = 0. (B) Color representation of response traces. Each line represents a single response trace. Responses to consecutive stimulation pulses, delivered at 15Hz, are ordered top to bottom. Voltage is color coded, red for high voltage and blue for low. It can be seen that spikes are fired in response to some of the stimulations, in a seemingly random and complex manner. (C) Examples for stimulation sequences, each of 1h length. The signals shown are stimulation rates, the reciprocals of the stimulation interval series. Examples are given of the three stimulation regimes (see Results section): constant interval (upper panel, blue), white noise (middle panel, red) and scale-free (bottom panel, green). (D) The autocorrelation function for the white noise (red) and scale free (green) sequences. (E) Power Spectral Density (PSD) for the white noise (red) and scale free (green) sequences. Figure 2. Response statistics under the three stimulation regimes. Main panels show raw data and statistical measures of a typical neuron to the three regimes (1h of stimulation for each sequence): constant interval (blue), white noise (red) and scale-free (green). Insets show results from a neuron subjected to a control experiment where the significance of the differences in responses to each of the regimes are assessed by 10 repetitions of each stimulation block. (A) Extracts from the firing rate of the neuron in response to the three stimulation types (8min length, 1s bin width). (B) Histograms of firing rate values from one repetition, calculated with 5s bin size. The inset shows the distribution of mean firing rate value from each of the 10 repetitions in the control experiment. Filled box represents the 25%-75% range, and whiskers extend to the extreme values. (C) Histograms of ISI values from one repetition. The inset plots the range of values of the ISI coefficient of variation under 10 repetitions. (D) Response PSD from one repetition, on double logarithmic axes. The inset depicts the range of exponent 18 values for a power law fit to the low frequency tail of the PSD. (E) Detrended Fluctuation Analysis (DFA, see Methods) of the responses from one repetition. The inset depicts the range of exponent values for a power law fit to the Fluctuation curve. (F) Allan variance of the responses of the responses from one repetition. The inset depicts the range of exponent values for a power law fit to the linear tail of the curves. Figure 3. Input and output correlations under the three stimulation regimes. (A) Extracts from the response rates of a neuron to long stimulation (1 h) under the three regimes. Responses and stimulation are binned with a 5s bin size, 30 bins are shown. Mean is subtracted to have the input and output aligned. The responses under white noise and scale-free regimes follow closely the stimulation (black line), while for constant interval the variability is freely running. (B) Extracts from the same experiment, using a 100s bin size, 30 bins are shown. The response in the scale-free regime is still locked to stimulus fluctuations, while input fluctuations in the white noise regime are substantially diminished, therefore unable to lock response variability. (C) A scatter plot of the stimulation rate against response rates, calculated with a 5s bin size. White noise input in red, scale-free input in green. Mean is removed for visual clarity, and both axes are in standardized units. While the marginal distributions for both inputs are similar, the correlation in the scale-free case is significantly higher. (D) Scaled correlation analysis. The correlation between input and output rates are calculated in different timescales, for white noise and scale-free regimes. The correlation in a given timescale T is calculated by smoothing the series with a rectangular window of width T , and subtracting a version of the series smoothed with a window of size 25T , effectively band-passing the time series. (E) Dependence of the input output correlation on the exponent β of the input. Neurons (n=31) were stimulated with blocks of 70 minutes duration with exponent ranging from 0 to 2. The input output correlation (cβ) was calculated from the responses of each block, using a 1s time bin. The graph shows the population statistics of the ratio cβ/c0 for each β, using box and whisker plot. The red horizon- tal line marks the median, the box marks the lower and upper quartiles and the whiskers the 19 range of data values. Outliers (values outside the range of 3 SD units from the average value) are marked with red points. (F) Distribution of preferred exponent values (the exponents which results in the strongest correlation for each neuron). (G) Extracts comparing the entrainment for β = 1 (left) and β = 2 (right), from a single neuron, demonstrating lack of entrainment for the fast fluctuations for the latter case. Traces are smoothed with a 2s rectangular window, time scale is identical for the two extracts. Figure 4. Repeatability of neuronal response. (A) The firing rates of a neuron under 10 identical stimulation sequences, under the three regimes. Responses to the constant interval stimulation show no reproducibility at all. Responses to the white noise input show repro- ducibility on short timescales, as can be seen in the inset. At longer timescales there is again no response repeatability. The responses to the scale-free sequence are the most reproducible, and lock to the input on many timescales. Stimulation rates are plotted in black for the three plots, but are normalized to have the mean and SD of the responses for clear visualization. (B) Pairwise distance matrices between responses to repetitions of the same stimulation sequence, calculated according to correlation metric (left) and the Victor-Purpura (VP) distance with tem- poral parameter q = 1 (right, see Methods section for details). Distance values are color coded, and for each metric the color scale was normalized to the maximal value. Diagonal pixels were whitened for visual clarity. (C) The purple line designates the mean and SD of the pairwise distances depicted in B, for the correlation metric, quantifying the change in repeatability under the three regimes. Also shown are the means for all 14 recorded neurons (grey for those which showed significant improvement for scale-free input regime, brown for those which didn't). (D) Same as C, but for VP distance. (E) Mean and SD of the average firing rate per trial for the neuron of A (purple). The firing rate significantly increases for scale-free input for this neuron. While this is not generally true across the population of neurons recorded, the firing rate never significantly decreases for scale-free input. (F) The effect of the temporal scale q on the results for the spike train metric. q is varied from 0.25s (light grey) to 10s (dark grey). While the 20 typical values differ, the effect exists for any choice of q, as can be expected from the lack of typical scale for the phenomena. The red curve is for q = 1. Figure 5. (A) Metric analysis on surrogate data. As explained in the Results section, the purple curve is metric analysis results, as in Figure 4, on surrogate data generated by applying a Bernoulli response model with a mean equaling the average response of the neuron, using VP distance with q = 1. The orange curve is results of the same analysis applied to data generated by a Bernoulli response model conditioned on the last interval, according to the curve in B. The grey curve is the actual experimental results for this neuron. (B) The response probability of a neuron, conditioned on the last stimulation interval, for white noise input (red, lower curve) and scale-free input (green, upper curve). Both curves, as expected, are mostly increasing. The difference between the curves implies the dependence of the probability on history longer than the last interval. Since the scale-free input contains correlations between intervals, its response curve differs from the white noise curve. (C) The effect of block length. The metric analysis of Figure 4 was repeated with various block sizes, i.e. with analysis performed on the first T seconds of each repetition, using the correlation metric. Data was taken from the onset of the stimulation block, including the transient phase. This analysis shows how the responses to the constant interval input and white noise input are drifting apart relatively rapidly (mean distance increases), while the responses to the scale-free input are forced together by the input dynamics, and show only a slow and moderate gain in distance. (D) Block length analysis as in C, using VP distance . It should be pointed that both the metrics used here are effectively normalized to the input length, in contrast to metrics like Euclidean and others which are extensive in input length. 21 Figure 1 22 51015−200−150−100−50050100time (ms)v (µV)interval −1 (1/s) −3003000.51time lag (s)corr0.010.11101001000frequency (Hz)psd3ms1minABCDE Figure 2 23 101001000110100bin size (s)F0.60.8051015sr (spk/s)Prob456012ISI (s)Log Prob1.61.822.2.001.01.11.1110100f (Hz)PSD−0.8−0.61101001000110100bin size (s)Allan Variance0.811.22minABCDEF Figure 3 24 −4−2024−4−2024stim ratespike rate05010000.20.40.60.8timescale (s)correlation0.250.751.251.750.60.811.21.41.61.82βcorrelation ratio0.250.751.251.750246810βcount10min30s1minABCDEFG Figure 4 25 cwsf0.51dcwsf0.1.2.3.4dcwsf468sr (spk/s)cwsf.1.3.5.7Dspt1min0100.4ABCDEF Figure 5 26 cwsf.15.20.25.30d0.1.2.3.6.9Interval (s)Spike Prob0120240360480600.2.4.6.81T (s)d0120240360480600.2.4T (s)dABCD
1301.5148
1
1301
2013-01-22T11:17:38
Limit-cycle-based control of the myogenic wingbeat rhythm in the fruit fly Drosophila
[ "q-bio.NC", "nlin.AO" ]
In many animals, rhythmic motor activity is governed by neural limit cycle oscillations under the control of sensory feedback. In the fruit fly Drosophila melanogaster, the wingbeat rhythm is generated myogenically by stretch-activated muscles and hence independently from direct neural input. In this study, we explored if generation and cycle-by-cycle control of Drosophila's wingbeat are functionally separated, or if the steering muscles instead couple into the myogenic rhythm as a weak forcing of a limit cycle oscillator. We behaviourally tested tethered flying flies for characteristic properties of limit cycle oscillators. To this end, we mechanically stimulated the fly's gyroscopic organs, the halteres, and determined the phase relationship between the wing motion and stimulus. The flies synchronized with the stimulus for specific ranges of stimulus amplitude and frequency, revealing the characteristic Arnold tongues of a forced limit cycle oscillator. Rapid periodic modulation of the wingbeat frequency prior to locking demonstrates the involvement of the fast steering muscles in the observed control of the wingbeat frequency. We propose that the mechanical forcing of a myogenic limit cycle oscillator permits flies to avoid the comparatively slow control based on a neural central pattern generator.
q-bio.NC
q-bio
Limit-cycle-based control of the myogenic wingbeat rhythm in the fruit fly Drosophila Jan Bartussek1, A. Kadir Mutlu1, Martin Zapotocky2,3, Steven N. Fry4 1 Institute of Neuroinformatics, University of Zurich and ETH Zurich, Winterthurerstrasse 190, 8057 Zurich, Switzerland 2 Institute of Physiology, Academy of Sciences of the Czech Republic, Videnska 1083, 14220 Prague 4, Czech Republic 3 Institute of Biophysics and Informatics, First Faculty of Medicine, Charles University, Salmovska 1, 12000 Prague 2, Czech Republic 4 SciTrackS GmbH, Lohzelgstrasse 7, 8118 Pfaffhausen, Switzerland Author Generated Postprint, published in J R Soc Interface 10:20121013. http://dx.doi.org/10.1098/rsif.2012.1013 Electronic Supplementary Material available under http://rsif.royalsocietypublishing.org/content/suppl/2012/12/28/rsif.2012.1013.DC1.htm l Author for correspondence: Jan Bartussek, e-mail: [email protected] Abstract In many animals, rhythmic motor activity is governed by neural limit cycle oscillations under the control of sensory feedback. In the fruit fly Drosophila melanogaster, the wingbeat rhythm is generated myogenically by stretch -activated muscles and hence independently from direct neural input. In this study, we explored if generation and cycle-by-cycle control of Drosophila’s wingbeat are functionally separated, or if the steering muscles instead couple into the myogenic rhythm a s a weak forcing of a limit cycle oscillator. We behaviourally tested tethered flying flies for characteristic properties of limi t cycle oscillators. To this end, we mechanically stimulated the fly’s ‘gyroscopic’ organs, the halteres, and determined the phase relationship between the wing motion and stimulus. The flies synchronized with the stimulus for specific ranges of stimulus amplitude and frequency, revealing the characteristic Arnol´d tongues of a forced limit cycle oscillator. Rapid periodic modulation of the wingbeat frequency prior to locking demonstrates the involvement of the fast steering muscles in the observed control of the wingbeat frequency. We propose that the mechanical forcing of a myogenic limit cycle oscillator permits flies to avoid the comparatively slow control based on a neural central pattern generator. 1. Introduction Locomotion plays a profound role in essential animal behaviours, such as foraging for food, finding mates and evading predation. In most cases, animals rely on the rhythmic actuation of appendages, such as legs, fins and wings for locomotion. While providing reaction forces for forward propulsion or to remain airborne, these same structures typically also serve as control surfaces to achieve stability of motion, as well as perform impressive manoeuvres. In many animals, the rhythmic activity underlying locomotion originates in neural central pattern generators (CPGs), whose cyclic activity is modulated by reafferent sensory feedback. CPGs represent a ubiquitous neural mechanism by which animals generate and control rhythmic activity for various bodily functions, such as respiration, chewing and limb actuation [1]. CPGs have been successfully modelled using oscillators with a stable limit cycle (limit cycle oscillators , LCOs) [2-4]. The intrins ic properties of such nonlinear oscillators are illustrated in Fig. 1 by example of the Van der Pol oscillator [5]. The oscillation of position in time (black trace in Fig. 1A) represents a limit cycle orbit in the position -velocity phase space (black curve in Fig. 1B). Limit cycle orbits are asymptotically stable, i.e. the system converges to the limit cycle following a transient perturbation. In the example shown, the delivery of pulse perturbations during three consecutive cycles (Fig. 1A) temporarily affects the periodic activity (green curve), which subsequently returns to its stable limit cycle (Fig. 1B). Depending on the cycle phase, in which such a forcing is applied, the frequency of the LCO is transiently affected [6]. In the present case, the delivery of the pulses leads to a frequency increase (compare green and dashed lines in Fig. 1 A). In such a way, precisely timed input can control the rhythmic pattern of LCOs efficiently and rapidly – a phenomenon that has also been experimentally demonstrated in CPGs [7, 8]. Computational modelling has shown that phase-dependent reafferent feedback from the periphery to the CPG can ensure robustness of the rhythm in presence of environmental perturbations [9-11]. Such limit-cycle-based control therefore provides a conceptual framework in which to understand how animals generate and control complex motor patterns robustly and efficiently [12]. Figure 1. A) Time course of a van der Pol oscillator perturbed by external force . The equation of motion is with . Blue trace: perturbation . Black trace: before and after perturbation. Green trace: during perturbation. Dashed trace: continuation of unperturbed oscillation (perturbation not switched on, ). B) Limit cycle of van der Pol oscillator. The black and green traces from A) are replotted in the phase space (velocity vs. coordinate ); increasing time corresponds to clockwise rotation. CPGs were originally discovered in locusts [13], in which CPGs provide the neural rhythm for the activity of the thoracic flight muscles giving rise to the wing motion. Synaptic input from the sensory system modulates these patterns to achieve flight control [14-16]. An entirely different actuation mechanism is found in flies, whose highly specialized flight apparatus likewise serves as a model to explore fundamental control mechanisms. The sophisticated flight control abilities of flies are exemplified by the tiny fruit fly Drosophila melanogaster , which performs sharp turning manoeuvres within a fraction of a second (around 50 milliseconds, or 10 wingbeats) [17]. Unlike in locusts, the fast and power demanding wingbeat of fruit flies arises myogenically, without direct neural control [18]: Antagonistic sets of stretch act ivated muscles bring the thorax and with it the wings into resonant oscillation at around 200 Hz. At the time scale of single wing strokes, the wing motion i s finely modulated by minuscule steering muscles, which insert directly at the sclerit es of the wing articulation and are under immediate neural control [19-22]. The activity of these direct steering muscles depends on a reflexive feedback loop involving specialized mechanosensory organs, the halteres, which sense the rotational velocity of the fly’s body [23-29]. According to the current understanding, the powerful, but “dumb” [30] stretch activated muscles supply a periodic force to drive the roughly sinusoidal wing motion [31, 32], while the weak, but fast direct steering muscles modulate stroke position and amplitude to stabilize flight and manoeuvre. In this way, the highly differentiated muscle types of flies are assumed to bring about the generation and control of the wingbeat rhythm as functionally separate processes [30, 33] - unlike in locusts, in which actuation and control are closely integrated within the framework of a LCO forced by afferent neural input. In the fly, it remains unclear if and how the myogenic rhythm, which cannot be controlled on a cycle-by-cycle basis through direct neuronal input, may be efficiently regulated by sensory feedback. While cycle-by-cycle control of wing stroke amplitude, stroke deviation and mean stroke position by the direct action of the steering muscles has been extensively studied [33-36], the fast control of the thorax/wing oscillator frequency has not been investigated in the previous literature. To achieve such frequency control, the myogenic rhythm would have to be coupled to the only system that is able to act on such fast time scales, i. e., the steering muscles. The null hypothesis of a functional separation between power and steering muscles predicts that flies are unable to control their wingbeat frequency at the time scale of single wing strokes. While flies apparently do not rely on CPGs to provide their wingbeat rhythm, it is intriguing to consider the possibility that the indirect muscle actuation mechanism itself represents a LCO, which is forced in a phase-dependent manner by the )(tx)(tF)()1(2tFxxxx3.0)(tF)(tx)(tx0)(tFdtdx/xt mechanical activity of the steering muscles. This alternative hypothesis is consistent with the “limit-cycle control” scheme for insect flight suggested in [37, 38]. If confirmed, the flight control strategy of flies, and the functional role of their steering muscles in particular, would appear in a new light altogether. We explored this possibility by testing fruit flies for a functional coupling of the steering muscles onto the wing stroke pattern generated by the stretch activated muscles. We evoked steering muscle activity from a periodic mechanic stim ulation of the halteres, which are known to provide direct input to the steering muscles [21]. We used a laser vibrometer to measure stroke-by-stroke variations in the wingbeat frequency of tethered flying flies and attribute these changes unambiguously to the activity of the direct steering muscles, which are the only muscles in the flight apparatus of flies that are capable of responding at the time scale of a single wing stroke. In the case of sustained periodic forcing, a LCO may adjust its frequency and eventually synchronize with the forcing, a phenomenon that has been observed in various biological limit cycle oscillators (e.g. [39, 40], review [41]). Depending on the value of the forcing frequency and amplitude, one can expect to obtain 1:1 synchronization (entrainment), or a regime in which the forcing frequency and the LCO frequency are other rational multiples of each other, such as 2:1, 3:2, etc.. Consistent with this generic property of periodically forced LCOs, we found 4 distinct regions of s timulus parameter space (so called Arnol’d tongues), in which the flies synchronized their wingbeat with the forcing. Furthermore, we found that the stimulation caused a fast modulation of the wingbeat frequency just outside of these parameter regions, demonstrating control of the wingbeat rhythm on a cycle-by-cycle basis. Our results indicate that the direct steering muscles of flies function as a weak mechanical forcing of a limit cycle oscillator embodied in the myogenic wing actuation mechanism. 2.1 Flies 2. Materials & Methods We obtained fruit flies (Drosophila melanogaster Meigen) from our laboratory stock (descended from a wild-caught population of 200 mated females). A standard breeding procedure was applied (25 females and 10 males, 12:12 hour light/dark cycle, standard nutritive medium). The experiments were performed during the first 8 hours of a subjective day with 5-10 days old female flies. 2.2 Experiments We cold- anaesthetized single fruit flies and glued them by their thorax to the tip a steel tether (Fig. 2A). We then attached the tether to a piezoelectrical actuator (P830.40, Physik Instrumente, Germany). We used a function generator (Agilent 33120A, Hewlett-Packard, USA) to create a sinusoidal voltage signal, amplified it using a custom built amplifier and applied to the piezo. The piezo-induced displacement of the tether was proportional to the voltage with a conversion factor of 0.9 µm/V. The resulting sinusoidal displacement of the tethered fly caused an inertial force Fi on the endknobs of its halteres, given by , with the amplitude of the tether displacement , the signal’s frequency and the mass of the haltere endknob [24]. We oscillated the flies along the anterior-posterior axis, applying either voltage frequency sweeps (0.1-500 Hz, sweep rate 5 Hz/s, holding constant the voltage at 1, 2, 4, 6, 8 or 10 volts) or voltage amplitude sweeps (0.1-8 V, sweep rate 0.8 V/s, holding constant the frequencies close to the baseline wingbeat frequency of the fly currently being tested). Note that as the inertial force is proportional to the square of the stimulus frequency, the forcing amplitude slowly increases during a voltage frequency sweep - despite the voltage amplitude at the piezo staying constant. When the stimulus frequency is equal to (or very near) the wingbeat frequency, and when the stimulus and the wingbeat are in the appropriate phase relation, the inertial force Fi resembles the Coriolis forces acting on the halteres during pitch manoeuvres. Specifically, to mimic the time course of the Coriolis forces, the phase relation must be such that the tether acceleration is zero at the time points corresponding to the dorsa l and ventral wing reversal [28]. The amplitude range of our stimulation (1 - 10 V) then corresponds to pitch rotations of 120 - 1200 °/s (see electronic supplementary material), which lies in the range typical for free flight manoeuvres [42]. We recorded the stimulation and the fly’s response together by measuring the velocity of the tether vibration about 1 mm above the fly’s body using a laser Doppler vibrometer (MSA-500, Polytec, Germany), sampling at 10240 Hz. We chose HTimtAtFsin)(2TAHm appropriate tethers (stainless steel, 1cm length, 200µm diameter) to tune the system suitably for our measurements. In each wing cycle, the wings generate a sharp force peak that is known to occur near the ventral reversal phase of the wingbeat [43]. This leads to an overdamped, high-frequency (>1000 Hz) oscillation of the tether within each wing cycle. In the measured tether velocity trace, this fly-induced signal was overlaid with a low frequency signal corresponding to the piezo actuator (0.1-500 Hz) – see Fig. 2B, top trace. Figure 2. A) Experimental setup. Flies were tethered to a steel pin, which was attached to a piezoelectric actuator. The piezo oscillated the tether, and with it the fly, according to the applied voltage s ignals created by a function generator. In addition, the fly’s wingbeat causes the tether to vibrate; we measured the total tether oscillation using a laser Doppler vibrometer. B) Data processing. Top panel: Raw data of the tether vibration generated by a flying fly and a sinusoidal piezo oscillation with a frequency close to the baseline WBF of the fly (about 200 Hz). Middle panel: The raw data was high-pass and low-pass filtered to separate the fly’s signal (spiky trace, red) from the piezo signal (sinusoidal trace, blue). Bottom panel: Instantaneous phase of the piezo oscillation (blue trace) and the phase in which the main peak in the fly’s signal occurred (red dots). C) Schematic Arnol’d tongues. Four selected regions of synchronization (filled triangles) for a generic limit cycle oscillator with natural frequency , forced by an external stimulus of frequency and amplitude . At low stimulus amplitudes, locking of the oscillator to the stimulus occurs when is close to a rational multiple of ; the frequency range increases with increasing . Arrows indicate amplitude and frequency sweeps (see text). 2.3 Data analysis Phase extraction To extract the wing stroke phase from the tether vibration, we applied a 7 th order high-pass Butterworth filter in a zero phase lag configuration with a cut-off frequency of 1000 Hz (red trace in Fig. 2B, middle panel). We then up-sampled the data to 50 kHz and applied a peak detection algorithm to determine the times of the prominent peak in this fly-induced s ignal. The times approximately coincide with the ventral reversal of the wings. To extract the stimulus phase, we applied a 7 th order low-pass Butterworth filter with a cut-off frequency of 500 Hz (blue (blue trace in Fig. trace in Fig. 2B, middle panel). We then applied the Hilbert transform to obtain the forcing phase 2B, bottom panel). Zero phase (i.e., = 0 modulo ) was chosen to coincide with the time points at which the tether natfstimfstimastimfnatfstimaktkt)(t2 moves with maximal velocity in the direction towards the posterior direction of the fly (Fig 2A). Note that when = 0 and , the acceleration of the tether becomes zero. Synchrogram construction We applied a method called synchrogram analysis to reveal synchronization between the stimulus and the response in our data. This method was developed and successfully applied to detect phase-locking in nonstationary and irregular signals [44, 45]. A synchrogram can be described as a ‘phase stroboscope’: Intervals of synchronization between an ex ternal forcing and an oscillator are revealed by plotting the phase of the oscillator at periodic instances of the forcing, or vice versa. We constructed the synchrograms by computing the phase of the forcing at the periodic instances of the wingbeat, (red dots, Fig. 2C). When the fly’s wing stroke and the applied force are not synchronized, changes rapidly (Fig. 2B, bottom panel, 0 - 0.7s). When the fly phase-locks to the stimulus and beats its wings at the f orcing frequency (1:1 synchronization), remains constant (single, almost horizontal line in Fig. 2B, bottom panel, 0.7 - 1s). In the case of higher-order synchronization (i.e. n wingbeats fit to m stimulus cycles , n:m locking), the phase points form m roughly horizontal lines in the synchrogram. Automated synchrogram evaluation To determine the stimulus parameter regions in which synchronization took place (i.e. the Arnol’d tongues), we adapted a synchronization detection algorithm for synchrograms from [46] and [47]. Because higher-order lockings tend to be difficult to detect [6], we restricted our analys is to four low-order lockings (1:2, 1:1, 3:2 and 2:1). The synchronizat ion detection algorithm operates as follows on the synchrogram time series . First, the phase is rewrapped according to the locking index m, such that modulo in the vicinity of each potential n:m locking region. Next, for each data point in the series, a window spanning 10 wingbeat cycles preceding it and following it is defined, and the mean and the standard deviation of in this window are calculated. If the standard deviation is higher than a threshold of 2.5% of , the data point is deleted. Following this step, the time series is defined on a single or multiple intervals in which the wingbeat is phase locked with the stimulus. Multiple intervals are present if a transient loss of locking occurs while still within the synchronization region (so called phase slips [6], see Sec. 3.1 for an example). To determine the width of the synchronization region, we therefore merged multiple intervals by linear interpolation if 1) their length was at least 15 poi nts (wingbeats) and 2) their distance was less than 10 points. The longest remaining interval was considered to be the candidate region of synchronization. To exclude false positives, i.e. intervals in which an increase in WBF was not caused by the stimulat ion but by random fluctuations in the wingbeat frequency, the candidate region was classified as “synchronized” only if the total number of points in the interval was above a threshold value . To determine , we constructed synchrograms from control data measured without piezo stimulation. This control synchrogram was constructed as described above, but in this case the measured phase of the stimulus was replaced by the phase of a fictitious stimulus. The synchrogram analys is was then run as described in the previous paragraph. We found that by setting for 1:2 locking, for 1:1 locking, for 3:2 locking and for 2:1 locking, our algorithm discarded at least 99.5 % of lockings (of each order) in the control synchrogram. Consequently, we used these threshold values of interval length to declare a candidate region as “synchronized” in our test data sets (with piezo stimulation). The first and the last point of the synchronized interval were taken as beginning and end of the respective Arnol’d tongue. Construction of Arnol’d tongues To place these border points into the frequency–amplitude parameter space of the forcing, we determined the stimulus amplitudes, i.e., the acceleration amplitude of the tether at the beginning and end of the synchronized interval. The acceleration was calculated from the measured velocity signal according to [48]. To account for possible differences in the baseline frequency (i.e., the wingbeat frequency in absence of piezo stimulation) in between tests and in the 6 flies tested; we divided the stimulus frequency at the border points by the fly’s mean wingbeat frequency in the 3 seconds preceding the synchronized interval plus the 3 seconds following the synchronized interval. For each frequency sweep test during which synchronization of a given order occurred, we therefore placed two border points into the parameter space, with the ratio of stimulus frequency to baseline frequency on the abscissa, and the acceleration delivered by the piezo on the ordinate. The set of all such border points (from the 6 flies tested) defines the Arnol’d tongu e. )(t)(t)(kt)(kt)(kt)(kt)()(kkmttm2mm2nmNnmN4512N14011N11732N21821N 3. Results 3.1 Nonlinear response to periodic mechanical stimulus When we stimulated flies periodically with the piezo, we observed nonlinear responses typical for forced limit cycle oscillators. We will first present three examples to describe these phenomena in detail. Periodically forced limit cycle oscillators are generically expected to exhibit multiple Arnol’d tongues, corresponding to synchronization at various commensurate ratios of stimulation frequency and oscillator frequency. The frequency range in which locking of a given order occurs typically increases with the forcing amplitude and gives rise to the characteristic tongue shape (Fig. 2C). A transition to synchronization should occur when the stimulus parameters are varied so that the border of the tongue is crossed. This can be achieved e.g. by a stimulus amplitude sweep (vertical arrow in Fig. 2C) or stimulus frequency sweep (horizontal arrow in Fig. 2C). To test for such a transition at the border of the 1:1 tongue (wingbeat frequency equal to stimulus frequency), we stimulate d flies with amplitude sweeps at fixed frequencies close to the fly’s baseline frequency. For example, we increased the amplitude from 0 to 3.8 m/s at a fixed detuning (difference of baseline frequency and stimulus frequency) of about 5 Hz (Fig.3A). Here, the synchrogram reveals synchronization as a line of constant relative phase (black dots, top trace), while the stimulus amplitude is above 2.3 m/s 2 (red dotted line, bottom trace). This phase locking lasts for about 7 seconds or about 1650 wingbeats. It is interrupted by three intervals, during which the phase changes rapidly (arrows in Fig. 3A). The occasional occurrence of such rapid phase changes, called phase slips, is typical for biological oscillators due to their inherent noisiness [6]. The calculated wingbeat frequency (WBF, black line, bottom trace) shows that the fly returns to its baseline frequency during the phase slips . Figure 3. A) Example of a recorded amplitude sweep. We stimulated with Hz while linearly first increasing and then decreasing , thus crossing the 1:1 Arnol’d tongue border vertically. The fly’s baseline WBF was 235 ± 3 Hz. Top: Relative phase between stimulation and wingbeat. Bottom: WBF, (dashed, blue), (dotted, red). Arrows on top indicate phase slips. B) Example of a recorded frequency sweep. We stimulated with mean m/s2 while linearly increasing , thus crossing the 1:1 Arnol’d tongue border horizontally. The fly’s baseline frequency was 211 ± 2 Hz. Top: Relative phase between stimulation and wingbeat. Bottom: WBF, (dashed, blue). Arrows indicate quasiperiodic modulation of the WBF. . Similarly, we tested for a transition to synchronization when decreasing the detuning. We stimulat ed flies with frequency sweeps with starting frequencies well below their WBFs . We then increased the stimulus frequency while keeping the applied voltage amplitude constant. In the example shown in figure 3B, the fly’s baseline WBF was 209 ± 4 Hz (mean ± std.) and we stimulated with a frequency sweep (0.1-500 Hz, 10 V). The fly locked to the forcing when the stimulation frequency was above 199 Hz and de-locked when it reached 220 Hz. The locking lasted for about 4 seconds or 835 wingbeats. Before and after looking, the WBF oscillated (see arrows). This oscillation reflects a regime in which the forcing frequency is too far away from fly’s baseline frequency to fully entrain the fly; however it is close enough to significantly influence the oscill ator. Indeed, similar frequency oscillations are generically expected for limit cycle oscillators when the forcing parameters are just 230stimfstima)(ktstimfstima1.8stimastimf)(ktstimf)(kmt outside the Arnol’d tongues (called phase walk-through or quasiperiodicity [6, 49]). The phenomenon is likewise visible in Fig. 3A just before the onset of locking. Besides 1:1 entrainment of the oscillator frequency to stimulus frequency, forced limit cycle oscillators typically also show higher-order synchronization (also called subharmonic and superharmonic entrainment in the literature). The corresponding Arnol’d tongues touch the abscissa axis (Fig. 2C) at points obtained as rational fractions of stimulus frequency and baseline frequency, fStim: fNat = n: m, where n, m are relatively prime integers larger than 1. The most robust tongues are generically expected for n, m small (for example 1:2, 3:2), while the tongues with large n, m tend to be very narrow and difficult to observe in the presence of noise. Figure 4. Synchrogram of a full frequency sweep. Relative phase between stimulation and wingbeat. We stimulated with a frequency sweep from 0.1 to 500 Hz with 10V applied piezo voltage. The baseline frequency of the fly was 209 ± 4Hz. The continuous recording is shown divided in three rows for clarity. Phase locking leads to the occurrence of horizontal or weakly tilted lines of phase points; for identified lockings of more than 100 wingbeats the corresponding Arnol’d tongue order n:m is indicated on top. Arrow indicates phase slip at 2:1 locking. To test for the existence of higher-order synchronization, we applied frequency sweeps and constructed synchrograms in a range from 40 to 455 Hz. As the baseline wingbeat frequencies were around 200 Hz, the sweeps were expected to cross the borders of multiple Arnol'd tongues. If, for example, a fly beating with 195 Hz is stimulated at 400 Hz, we expect the fly to increase its frequency to 200 Hz, so that the frequency ratio adjusts to the locking order 2:1. An example synchrogram of a full recording is shown in Fig. 4. Here, the synchrogram reveals not only the extended 1:1 locking region (around 210 Hz), but also several lockings of higher order n:m. The fly synchronized in intervals in which the phase points form a single horizontal line (for locking orders with m=1, like 1:1 and 2:1) or m horizontal lines (for m>1, see also Sec. 2.2). In the presented example, all prominent lockings with a minimum length of 100 wingbeats are marked on top with the corresponding locking order. Note the phase slip at the beginning of the 2:1 locking. This measurement contains 4 prominent lockings, with 3:2 being the highest locking order. 3.2 Locking statistics and Arnol’d tongues To extract the Arnol’d tongue boundaries and to quantify the occurrence of phase-locking, we focused on the frequency sweep experiments. This had the major advantage that by applying a single sweep, we crossed the borders of multiple Arnol'd tongues and, in case of synchronization, obtained both an entry and an exit point of the respective Arnol ’d tongue. We tested )(kt 6 flies and stimulated each with 6 frequency sweeps of varying voltage amplitude (see Sec. 2.3). We obtained results from 33 measurements (Table 1); during 3 trials, the flies stopped flying and the respective measurements were discard ed. 1:1 locking occurred in 16 out of 33 trials (48%). 1:2, 3:2 and 2:1 lockings occurred in 56%, 52% and 33% cases, respectively. We found 1:1 entrainment and higher-order locking in every fly. The percentage of sweeps, in which locking was observed in a fly, was 47 ± 10% (mean ± std.). Figure 5. Extracted Arnol’d tongues. Start (>) and end points (<) of synchronization as automatically detected from the frequency sweep synchrograms. For the 1:1 lockings we show linear fits of the boundaries of the tongue (robust fits, forced through [1,0]). From the control experiments with flies whose haltere endknobs were ablated, we obtained only two lockings (►◄) with markedly reduced width. From the identified locking intervals, we reconstructed the Arnol’d tongue boundaries, as described in Sec. 2.3. We found that the 1:1 tongue is roughly triangular, i.e. an increasing forcing strength leads to a roughly proportional increase in locking width, as expected for a weakly forced limit cycle oscillator (Fig. 4, second panel). The width of the tongue shows that the synchronization is maintained for a detuning up to ± 5% of the WBF (which corresponds to about ±1 0 Hz). The shapes of the higher-order tongues (Fig. 5, panels 1, 3 and 4) are more inconclusive and do not appear to be triangular. Still, all tongues cover extended areas in the frequency-amplitude space. Note that each measured tongue is nearly symmetrical, which is a generic property of very weakly forced limit cycle oscillators [6]. While it cannot be entirely excluded that a hysteretic effect was nevertheless present due to a frequency scanning in one direction only, it is rather unlikely that this would have exactly compensated an existing intrinsic asymmetric shape. Furthermore, the possible existence of a hysteresis effect does not challenge any of the drawn conclusions in this work. Table 1. Locking quantification. Lockings rates, i.e. percentage of frequency sweeps in which phase locking was observed, obtained from 33 frequency sweeps at piezo voltage amplitudes of 1-10 V. Fly 1 Fly 2 Fly 3 Fly 4 Fly 5 Fly 6 Locking rate per locking order Locking rate per 1:2 100 25 67 50 60 33 1:1 100 50 67 33 20 17 3:2 2:1 83 50 67 33 60 17 83 25 67 0 20 0 fly 92 38 67 29 40 17 Mean locking rate per locking order 56 48 52 33 rate per fly ± std. Mean locking 47±10 3.3 Fast and phase-dependent modulation of wingbeat frequency We showed that sustained mechanical stimulation at a fixed frequency (or slowly increasing frequency in the case of a frequency sweep) acts as a forcing that can entrain the wingbeat rhythm. This result, however, does not tell us the time scal e on which the forcing acts. To deduce this time scale, we analysed the intervals of fast WBF modulation prior to 1:1 locking (see Sec. 3.1) in more detail. Such intervals were readily identified for stimulus amplitudes of more than 2 m/s 2. In these intervals, the WBF was found to oscillate with a frequency that increased with the difference between baseline WBF and stimulation frequency. We observed WBF oscillations with frequencies up to 40 Hz. In the example shown in Fig. 6A, the time courses of WBF and the phase of stimulation reveal that the WBF varies from cycle to cycle, and follows the phase of stimulus relative to the wingbeat. For further analys is, we selected recordings, in which the WBF oscillation lasted for at least 2 seconds (7 measurements from 4 flies). The range of modulation frequencies was 11 to 18 Hz (mean 14Hz, std. 2.5 Hz). We found that the response to stimulus is cons istent across the recordings, and that the resulting change in WBF varies approximately sinusoidally wit h the ) can be well fitted (r2 = 0.92) with the first two components of its Fourier series (Fig. 6B). A maximal decrease in WBF is observed around relative (Fig. 6B). The observed response funct ion (which by definition is periodic in relative phase phase , while a phase offset of 0 led to a maximal increase in WBF. Stimulation with phase offsets of and did not elicit a response in WBF. Recall that when the stimulus is applied with the frequency close to WBF and with the relative phase or , the resulting acceleration has a time course mimicking the Coriolis force acting on the haltere knobs during a pitch manoeuvre in free flight (see Sec. 2.3). For other values of , the time course of the acceleration does not match the Coriolis force in any rotational manoeuvre during free flight. The response of the wing/thorax oscillator is thus maximal when the stimulus mimics the Coriolis force, to which the haltere mechanosensory pathway is known to be highly sensitive [25]. Figure 6. A) Example of fast modulation of wingbeat frequency during stimulus frequency sweep. Recording of 1 second of flight prior to the onset of 1:1 locking. Upper trace: Relative phase between stimulation and wingbeat. Lower trace: WBF. B) Phase- coupled frequency response. Pooled results from 7 recordings of stimulus frequency sweeps in 4 flies. For each recording, we calculated the difference instantaneous WBF and average WBF over the interval [ -1.5s:-0.5s] prior to locking, and normalized the time series by its variance. We then binned the data with respect to the stimulus phase. Pooling the responses from the 7 rec ordings, we calculated the median (open circle) and the interquartile range (error bars). Trace: best fit for the two leading Fourier components of the periodic response function. )(ktf)1.32sin(06.0)52.1sin(55.0xxy2/2/0 Based on the observed phase-dependent response, we can make a deduction on how quickly the stimulus affects the WBF. Suppose that a fixed time delay elapses between the mechanical stimulus and the resulting WBF adjustment. In the response curve in Fig. 6B, the time delay would appear as a phase delay , where is the period of the modulation of WBF. The response curve was obtained, however, from data segments in which the modulation frequency varied widely (by 60%). Therefore a sizeable time delay would result in a range of phase delays, and after averaging, a clear phase-dependent response (as in Fig. 6B) would likely not be obtained. In addition, the coincidence of the optimal stimulus (resulting in maximal response) with the Coriolis force would be lost. This suggests that the time delay is small compared to the typical period of WBF modulation (about 0.07 sec, or 14 wingbeat cycles). Our analysis therefore indicates that the mechanical stimulus affects the WBF within a few wingbeat cycles. 3.4 Control experiments To test if the observed entrainment was indeed mediated by the halteres, we measured the responses of flies whose haltere endknobs had been ablated (which reduce the haltere sensitivity by about 90% without damaging the sensory fields at the haltere base [25]). We performed 6 frequency sweeps per fly. In 4 out of the 6 flies with ablated halteres, we found no entrainment. In two flies we observed a single, reduced locking region at the highest stimulation strengths (Fig. 5, filled triangles). Our direct observations of the wingbeat of tethered flying flies revealed no gross difference between flies with ablated halteres and untreated flies. Likewise, no qualitative differences were found in the vibrometry traces . Comparison of wingbeat frequencies (extracted from the tether vibration) showed, however, an increased mean baseline WBF (225 ± 14 Hz) in flies with ablated halteres, which was about 12 % elevated, compared to untreated flies (202 ± 16 Hz). A similar observation was made by Dickinson [25], who reported an increase in WBF of about 24% following endknob ablation. 4. Discussion & Conclusions In this study, we explore the possibility that the steering muscles of flies affect the myogenic wingbeat rhythm as a mechanical forcing of a limit cycle oscillator (LCO). Our synchrogram analysis on the phase data of a mechanical stimulus and the wing stroke reveals several characteristic features of a LCO, namely phase-locking, phase-slips, higher-order locking and quasiperiodicity (Sec. 3.1, 3.2). We interpret the measured fast, phase-dependent control of the wingbeat frequency (Sec. 3.3) to result from the activity of the direct steering muscles that function as a mechanical forcing of a LCO embodied in the myogenic wing actuation mechanism. Our data refute the null hypothesis according to which the power-muscle generated rhythm is functionally decoupled from the activity of the direct steering muscles. We adopted our experimental approach from the seminal studies performed by Nalbach, who vibrated blowflies back -and- forth to emulate the Coriolis forces acting on the halteres of flies during rotations of the body in free flight [24, 28, 50, 51]. While these studies already revealed the basic synchronization phenomenon (1:1 locking, [50, 51])), they did not systematically explore the possible LCO properties of the flight apparatus. We extended the previous approach by sampling at a high rate the phase relationship between the wing stroke and the mechanical stimulus, using laser vibrometry. Furthermore, we systematically varied the frequency and amplitude of the applied mechanical stimulus to determine the Arnol’d tongues (i.e., regions of synchronization) of the flight control system, and to induce rapid modulation of wingbeat frequency prior to entrainment. Our results show that the fly responds to mechanical stimulation as a nonlinear oscillator. The essentially nonlinear prope rties we observed included transitions to / from synchronization, subharmonic / superharmonic entrainment, and phase slips. More specifically, our measurements showed behaviour typical for a nonlinear oscillator with a stable limit cycle [4, 6]. A special type of chaotic oscillator (with narrow strange attractor and a high degree of phase coherence) could have similar synchronization properties (see, e.g., [6], Sec. 10.1.2). Our observations are, however, most straightforwardly understood in the framework of a periodically forced LCO, and we have found no signs of an attractor other than a single limit cycle. Nalbach interpreted the 1:1 entrainment observed in blowflies as arising from a pitch illusion reflex [50]. Our experimental approach, however, revealed synchronization also at higher frequency ratios (1:2 and 3:2), at which the mechanical stimuli can no longer mimic Coriolis forces [28]. Hence, we interpret the observed synchronization as an emergent property of a periodically forced LCO embodied in the myogenic wing actuation mechanism of flies. T/T Our results are highly consistent with the notion that fast sensory input from the halteres mediates the forcing. First, the maximal response was obtained when the stimulus mimicked the time course of the Cori olis force that would act on the haltere knobs during a pitch manoeuvre (Sec. 3.3). This agrees with previous studies in which blowflies [24, 50] and fruit flies [25] were mechanically stimulated. Second, the responses were strongly reduced after ablating the end-knobs of the halteres (see control experiments described in Sec. 3.4). In our study, we used observations of rapid phase-dependent modulation of the wingbeat frequency prior to locking to infer that the forcing acts on a time scale comparable to one wingbeat cycle (Sec. 3.3). We attribute the periodic forcing to the mechanical action of direct steering muscles, which have the fast response properties that can explain the measured strict phase relationship between the mechanical stimulus and the fr equency response (Fig. 6). The basalar muscle M.b1 could serve this function, as it is known to modulate the wing kinematics according to the phase of its single contractions during successive wing strokes and receives fast monosynaptic input from the halt eres [21, 22, 33, 34]. It might be argued that the halteres directly affect the thoracic rhythm mechanically, rather than acting through the mechanosensory system and steering muscles. We evaluate this possibility as follows. We first consider the order of magnitudes of forces exerted on the thorax by the halteres. During 1:1 locking in our experiments, the acceleration reaction force due to the piezo actuation peaks at around N (see electronic supplementary material). We then compar e it with the estimated force exerted by a steering muscle. In the blowfly Calliphora, the peak force of the steering muscle M.b1 is of order 10-2 N and depends substantially on the phase in which it is activated (Fig. 8 in [52]). Given that Drosophila is 10 times smaller in length than Calliphora and following the scaling laws for insect flight muscles [53], we estimate the forces generated by the homolog M.b1 muscle in Drosophila to be 103 = 1000 times smaller than in Calliphora. Hence, the forces generated by Drosophila‘s steering muscles (10 µN) exceed the reaction acceleration forces from the halteres (60 nN) by more than 2 orders of magnitude. We therefore conclude that the mechanical forcing of the thoracic rhythm is predominantly effected by the steering muscles. While the limit cycle properties of central pattern generators (CPGs) are well understood, these are now also revealed in the myogenic wing actuation mechanism of an insect. It is per se not very surprising that a myogenic oscillation can be described as a limit cycle. Our study, however, reveals a coupling of the fly’s haltere mechanosensory pathway to the thorax/wing LCO that is functionally significant for flight control. This is evidenced by substantial changes in wingbeat frequency (abou t ± 10 Hz) measured when the halteres were stimulated by forces in the range of those occurring during typical flight manoeuvres. We concluded that this coupling is effected by the mechanical activity of the steering muscles. In this way , the flight apparatus of the fly is able to avoid the computationally expensive (and comparatively sluggish) neural mechanism of CPGs, and instead replace it with a direct realization of a mechanical limit cycle oscillator. A mechanical system of this type has been recently implemented in a swimming robot by Seo, Chung and Slotine [54]. Myogenic limit cycle – based control provides an elegant conceptual framework in which to understand how flies can achieve extremely fast and precise flight control with minimal neural resources. The fruit fly offers itself as a model system to explore limit cycle – based control mechanisms that minimize the required computational power and operate at high frequencies. Such knowledge could serve the design of biomimetic micro air vehicles (MAVs), which like flies are under severe constraints in terms of mass and power consumption, and therefore depend on highly efficient control strategies [55, 56]. Acknowledgements This work was supported by the Volkswagen foundation (I/80 984 - 986), Swiss National Science Foundation (CR23I2- 130111 / 1) and the Czech Republic (AV0Z50110509 and P304/12/G069). We thank Jonathon Howard, Henri Saleh, Vasco Medici and Hannah Haberkern for valuable discussions and support and an anonymous reviewer for insightful critical comments.. References 1 2 3 4 Delcomyn, F. 1980 Neural basis of rhythmic behavior in animals. Science 210, 492-498. (doi:10.1126/science.7423199) Cohen, A. H., Holmes, P. J., Rand, R. H. 1982 The nature of the coupling between segmental oscillators of the lamprey spinal generator for locomotion: A mathematical model. J. Math. Biol. 13, 345-369. (doi:10.1007/bf00276069) Collins, J., Richmond, S. 1994 Hard-wired central pattern generators for quadrupedal locomotion. Biol. Cybern. 71, 375-385. (doi:10.1007/bf00198915) Glass, L., Mackey, M. C. From clocks to chaos: The rhythms of life. 1988. 8106 5 6 7 8 9 10 11 van der Pol, B. 1927 On relaxation-oscillations. The London, Edinburgh and Dublin Phil. Mag. & J. of Sci. 2, 978-992. Pikovsky, A., Rosenblum, M. & Kurths, J. 2001 Synchronization - A universal concept in nonlinear sciences . Cambridge: Cambridge University Press. Vogelstein, R. J., Etienne-Cummings, R., Thakor, N. V., Cohen, A. H. 2006 Phase-dependent effects of spinal cord stimulation on locomotor activity. Neural Systems and Rehabilitation Engineering, IEEE Trans. 14, 257-265. (doi:10.1109/tnsre.2006.881586) Vogelstein, R., Tenore, F., Etienne-Cummings, R., Lewis, M., Cohen, A. 2006 Dynamic control of the central pattern generator for locomotion. Biol. Cybern. 95, 555-566. (doi:10.1007/s00422-006-0119-z) Taga, G., Yamaguchi, Y., Shimizu, H. 1991 Self-organized control of bipedal locomotion by neural oscillators in unpredictable environment. Biol. Cybern. 65, 147-159. (doi:10.1007/bf00198086) Kuo, A. D. 2002 The relative roles of feedforward and feedback in the control of rhythmic movements. Motor Control 6, 129-145. Laszlo, J., van de Panne, M., Fiume, E. Limit cycle control and its application to the animation of balancing and walking. Proc . 23rd annual conf. on Computer graphics and interactive techniques. 237231: ACM; 1996. p. 155-162. 12 Ijspeert, A. J., Crespi, A., Ryczko, D., Cabelguen, J.-M. 2007 From swimming to walking with a salamander robot driven by a spinal cord model. Science 315, 1416-1420. (doi:10.1126/science.1138353) Wilson, D. M. 1961 The central nervous control of flight in a locust. J. Exp. Biol. 38, 471-490. Wendler, G. 1974 The influence of proprioceptive feedback on locust flight co-ordination. J. Comp. Physiol. A 88, 173-200. 13 14 (doi:10.1007/bf00695406) 15 Ausborn, J., Stein, W., Wolf, H. 2007 Frequency control of motor patterning by negative sensory feedback. J. Neurosci. 27, 9319- 9328. (doi:10.1523/jneurosci.0907 -07.2007) 16 Schmeling, F., Stange, G., Homberg, U. 2010 Synchronization of wing beat cycle of the desert locust, Schistocerca gregaria, by periodic light flashes. J. Comp. Physiol. A 196, 199-211. (doi:10.1007/s00359-010-0505-9) 17 Fry, S. N., Sayaman, R., Dickinson, M. H. 2003 The aerodynamics of free-flight maneuvers in Drosophila. Science 300, 495-498- -. (doi:10.1126/science.1081944) 18 19 Pringle, J. W. S. 1949 The excitation and contraction of the flight muscles of insects. J. Physiol. (Lond.) 108, 226-232. Heide, G. 1974 The influence of wingbeat synchronous feedback on the motor output systems in flies. Z. Naturforschung C C 29, 739-744. 20 Heide, G. 1983 Neural mechanisms of flight control in Diptera. In Biona- report 2 (ed. W. Nachtigall), pp. 35 -52, Stuttgart, New York: G. Fischer Verlag. 21 Fayyazuddin, A., Dickinson, M. H. 1996 Haltere afferents provide direct, electrotonic input to a steering motor neuron in the blowfly, Calliphora. J. Neurosci. 16, 5225-5232. 22 Fayyazuddin, A., Dickinson, M. H. 1999 Convergent mechanosensory input structures the firing phase of a steering motor neuron in the blowfly, Calliphora. J. Neurophysiol. 82, 1916-1926. 23 Fox, J. & Daniel, T. 2008 A neural basis for gyroscopic force measurement in the halteres of Holorusia. J. Comp. Physiol. A 194, 887-897. (doi:10.1007/s00359-008-0361-z) 24 Nalbach, G., Hengstenberg, R. 1994 The halteres of the blowfly calliphora: I I. Three-dimensional organization of compensatory reactions to real and simulated rotations. J. Comp. Physiol. A 175, 695-708. (doi:10.1007/BF00191842) 25 Dickinson, M. H. 1999 Haltere-mediated equilibrium reflexes of the fruit fly, Drosophila melanogaster. Phil. Trans. R. Soc. B 354, 903-916. 26 Fraenkel, G. & Pringle, J. W. S. 1938 Halteres of flies as gyroscopic organs of equilibrium. Nature 141, 919-920. (doi:10.1038/141919a0) 27 28 Pringle, J. W. S. 1948 The gyroscopic mechanism of the halteres of Diptera. Phil. Trans. R. Soc. B 233, 347-384. Nalbach, G. 1993 The halteres of the blowfly Calliphora: I. Kinematics and dynamics. J. Comp. Physiol. A 173, 293-300. (doi:10.1007/BF00212693) 29 Nalbach, G. 1994 Extremely non-orthogonal axes in a sense organ for rotation: Behavioural analysis of the Dipteran haltere system. Neurosc. 61, 149-163. (doi:10.1016/0306-4522(94)90068-X) 30 31 32 Dickinson, M. H., Tu, M. S. 1997 The function of Dipteran flight muscle. Comp. Biochem. Phys. A 116, 223-238. Greenewalt, C. H. 1960 The wings of insects and birds as mechanical oscillators. Proc. Am. Phil. Soc. 104, 605-611. Ellington, C. P. 1984 The aerodynamics of hovering insect flight. III. Kinematics. Phil. Trans. R. Soc. B 305, 41. (doi:citeulike- article-id:8698273) 33 Heide, G., Götz, K. G. 1996 Optomotor control of course and altitude in Drosophila melanogaster is correlated with distinct activities of at least three pairs of flight steering muscles. J. Exp. Biol. 199, 1711-1726. 34 Tu, M. S., Dickinson, M. H. 1996 The control of wing kinematics by two steering muscles of the blowfly (Calliphora vicina). J. Comp. Physiol. A 178, 813-830. (doi:10.1007/BF00225830) 35 Balint, C. N., Dickinson, M. H. 2001 The correlation between wing kinematics and steering muscle activity in the blowfly Calliphora vicina. J. Exp. Biol. 204, 4213-4226. 36 Balint, C. N., Dickinson, M. H. 2004 Neuromuscular control of aerodynamic forces and moments in the blowfly, Calliphora vicina. J Exp Biol. 207, 3813-3838. (doi:10.1242/jeb.01229) 37 Taylor, G. K., Zbikowski, R. 2005 Nonlinear time-periodic models of the longitudinal flight dynamics of desert locusts Schistocerca gregaria. J. R. Soc. Interface 2, 197-221. 38 Żbikowski, R., et al. 2006 On mathematical modelling of insect flight dynamics in the context of micro air vehicles. Bioinspir. Biomim. 1, R26. (doi:10.1088/1748-3182/1/2/R02) 39 Stoop, R., Schindler, K., Bunimovich, L. A. 2000 When pyramidal neurons lock, when they respond chaotically, and when they like to synchronize. Neurosc. Res. 36, 81-91. (doi:10.1016/s0168-0102(99)00108-x) 40 Buck, J., Buck, E., Case, J. F., Hanson, F. E. 1981 Control of flashing in fireflies. J. Comp. Physiol. A 144, 287-298. (doi:10.1007/BF00612560) 41 42 Glass, L. 2001 Synchronization and rhythmic processes in phys iology. Nature 410, 277-284. (doi::10.1038/35065745) Dickson, W. B., Straw, A. D., Dickinson, M. H. 2008 Integrative model of Drosophila flight. AIAA Journal 46, 14. (doi:10.2514/1.29862) 43 Dickinson, M. H., Götz, K. G. 1996 The wake dynamics and flight forces of the fruit fly Drosophila melanogaster. J. Exp. Biol. 199, 2085-2104. 44 Schäfer, C., Rosenblum, M. G., Kurths, J., Abel, H.-H. 1998 Heartbeat synchronized with ventilation. Nature 392, 239-240. (doi:10.1038/32567) 45 Schäfer, C., Rosenblum, M. G., Abel, H.-H., Kurths, J. 1999 Synchronization in the human cardiorespiratory system. Phys. Rev. E 60, 857-870. 46 Bartsch, R., Kantelhardt, J. W., Penzel, T., Havlin, S. 2007 Experimental evidence for phase synchronization transitions in the human cardiorespiratory system. Phys. Rev. Lett. 98, 054102. (doi:10.1103/PhysRevLett.98.054102) 47 Hamann, C., Bartsch, R. P., Schumann, A. Y., Penzel, T., Havlin, S., Kantelhardt, J. W. 2009 Automated synchrogram analysis applied to heartbeat and reconstructed respiration. Chaos 19, 015106. (doi:10.1063/1.3096415) 48 Woltring, H. J. 1985 On optimal smoothing and derivative estimation from noisy displacement data in biomechanics. Hum. Movement. Sci. 4, 229-245. (doi:10.1016/0167-9457(85)90004-1) 49 Ermentrout, G. B., Rinzel, J. 1984 Beyond a pacemaker's entrainment limit: Phase walk-through. Am. J. Physiol. Regul. Integr. Comp. Physiol. 246, R102-106. 50 Nalbach, G. Verhaltensuntersuchungen zur Funktion der Halteren bei der Schmeissfliege Calliphora erythrocephala mit echten und simulierten Drehreizen. PhD- Thesis, Eberhard-Karls-Universität Tübingen; 1991. 51 Nalbach, G 1988. Linear oscillations elicit haltere mediated turning illusions and entrainment in the blowfly Calliphora. In Proc.16th Neurobiol. Conf. 1988; Göttingen. 52 Tu, M. S., Dickinson, M. H. 1994 Modulation of negative work output from a steering muscle of the blowfly Calliphora vicina. J. Exp. Biol.192, 207-224. 53 Schilder, R. J., Marden, J. H. 2004 A hierarchical analysis of the scaling of force and power production by dragonfly flight motors. J. Exp. Biol.207, 767-776.(doi:10.1242/jeb.00817) 54 Seo, K., Chung, S.-J., Slotine, J.-J. 2010 CPG-based control of a turtle-like underwater vehicle. Auton. Robot. 28, 247-269. (doi:10.1007/s10514-009-9169-0) 55 Xinyan, D., Schenato, L. & Sastry, S. S. 2003 Model identification and attitude control for a micromechanical flying insect including thorax and sensor models. In Proc. IEEE Intern. Conf. on Robotics and Automation, ICRA '03. (doi: 10.1109/ROBOT.2003.1241748) 56 Chung S.-J.; Dorothy M. 2010 Neurobiologically inspired control of engineered flapping flight. J. Guid. Control Dynam. 33, 440- 453. (doi:10.2514/1.45311)
1611.00388
2
1611
2017-01-22T01:48:48
Automated scalable segmentation of neurons from multispectral images
[ "q-bio.NC", "q-bio.SC" ]
Reconstruction of neuroanatomy is a fundamental problem in neuroscience. Stochastic expression of colors in individual cells is a promising tool, although its use in the nervous system has been limited due to various sources of variability in expression. Moreover, the intermingled anatomy of neuronal trees is challenging for existing segmentation algorithms. Here, we propose a method to automate the segmentation of neurons in such (potentially pseudo-colored) images. The method uses spatio-color relations between the voxels, generates supervoxels to reduce the problem size by four orders of magnitude before the final segmentation, and is parallelizable over the supervoxels. To quantify performance and gain insight, we generate simulated images, where the noise level and characteristics, the density of expression, and the number of fluorophore types are variable. We also present segmentations of real Brainbow images of the mouse hippocampus, which reveal many of the dendritic segments.
q-bio.NC
q-bio
Automated scalable segmentation of neurons from multispectral images Uygar Sümbül Grossman Center for the Statistics of Mind and Dept. of Statistics, Columbia University Douglas Roossien Jr. University of Michigan Medical School Fei Chen MIT Media Lab and McGovern Institute Nicholas Barry MIT Media Lab and McGovern Institute Edward S. Boyden MIT Media Lab and McGovern Institute Dawen Cai University of Michigan Medical School John P. Cunningham Grossman Center for the Statistics of Mind and Dept. of Statistics, Columbia University Liam Paninski Grossman Center for the Statistics of Mind and Dept. of Statistics, Columbia University Abstract Reconstruction of neuroanatomy is a fundamental problem in neuroscience. Stochastic expression of colors in individual cells is a promising tool, although its use in the nervous system has been limited due to various sources of variability in expression. Moreover, the intermingled anatomy of neuronal trees is challenging for existing segmentation algorithms. Here, we propose a method to automate the segmentation of neurons in such (potentially pseudo-colored) images. The method uses spatio-color relations between the voxels, generates supervoxels to reduce the problem size by four orders of magnitude before the final segmentation, and is parallelizable over the supervoxels. To quantify performance and gain insight, we generate simulated images, where the noise level and characteristics, the density of expression, and the number of fluorophore types are variable. We also present segmentations of real Brainbow images of the mouse hippocampus, which reveal many of the dendritic segments. 1 Introduction Studying the anatomy of individual neurons and the circuits they form is a classical approach to understanding how nervous systems function since Ramón y Cajal's founding work. Despite a century of research, the problem remains open due to a lack of technological tools: mapping neuronal structures requires a large field of view, a high resolution, a robust labeling technique, and computational methods to sort the data. Stochastic labeling methods have been developed to endow individual neurons with color tags [1, 2]. This approach to neural circuit mapping can utilize the light microscope, provides a high-throughput and the potential to monitor the circuits over time, and complements the dense, small scale connectomic studies using electron microscopy [3] with its large field-of-view. However, its use has been limited due to its reliance on manual segmentation. The initial stochastic, spectral labeling (Brainbow) method had a number of limitations for neuro- science applications including incomplete filling of neuronal arbors, disproportionate expression of the nonrecombined fluorescent proteins in the transgene, suboptimal fluorescence intensity, and color shift during imaging. Many of these limitations have since improved [4] and developments in 30th Conference on Neural Information Processing Systems (NIPS 2016), Barcelona, Spain. arXiv:1611.00388v2 [q-bio.NC] 22 Jan 2017 various aspects of light microscopy provide further opportunities [5–8]. Moreover, recent approaches promise a dramatic increase in the number of (pseudo) color sources [9–11]. Taken together, these ad- vances have made light microscopy a much more powerful tool for neuroanatomy and connectomics. However, existing automated segmentation methods are inadequate due to the spatio-color nature of the problem, the size of the images, and the complicated anatomy of neuronal arbors. Scalable methods that take into account the high-dimensional nature of the problem are needed. Here, we propose a series of operations to segment 3-D images of stochastically tagged nervous tissues. Fundamentally, the computational problem arises due to insufficient color consistency within individual cells, and the voxels occupied by more than one neuron. We denoise the image stack through collaborative filtering [12], and obtain a supervoxel representation that reduces the problem size by four orders of magnitude. We consider the segmentation of neurons as a graph segmentation problem [13], where the nodes are the supervoxels. Spatial discontinuities and color inhomogeneities within segmented neurons are penalized using this graph representation. While we concentrate on neuron segmentation in this paper, our method should be equally applicable to the segmentation of other cell classes such as glia. To study various aspects of stochastic multispectral labeling, we present a basic simulation algorithm that starts from actual single neuron reconstructions. We apply our method on such simulated images of retinal ganglion cells, and on two different real Brainbow images of hippocampal neurons, where one dataset is obtained by expansion microscopy [5]. 2 Methods Successful segmentations of color-coded neural images should consider both the connected nature of neuronal anatomy and the color consistency of the Brainbow construct. However, the size and the noise level of the problem prohibit a voxel-level approach (Fig. 1). Methods that are popular in hyperspectral imaging applications, such as nonnegative matrix factorization [14], are not immediately suitable either because the number of color channels are too few and it is not easy to model neuronal anatomy within these frameworks. Therefore, we develop (i) a supervoxelization strategy, (ii) explicitly define graph representations on the set of supervoxels, and (iii) design the edge weights to capture the spatio-color relations (Fig. 2a). 2.1 Denoising the image stack Voxel colors within a neurite can drift along the neurite, exhibit high frequency variations, and differ between the membrane and the cytoplasm when the expressed fluorescent protein is membrane- binding (Fig. 1). Collaborative filtering generates an extra dimension consisting of similar patches within the stack, and applies filtering in this extra dimension rather than the physical dimensions. We use the BM4D denoiser [12] on individual channels of the datasets, assuming that the noise is Gaussian. Figure 2 demonstrates that the boundaries are preserved in the denoised image. 2.2 Dimensionality reduction We make two basic observations to reduce the size of the dataset: (i) Voxels expressing fluorescent proteins form the foreground, and the dark voxels form the much larger background in typical Brainbow settings. (ii) The basic promise of Brainbow suggests that nearby voxels within a neurite have very similar colors. Hence, after denoising, there must be many topologically connected voxel sets that also have consistent colors. The watershed transform [15] considers its input as a topographic map and identifies regions associated with local minima ("catchment basins" in a flooding interpretation of the topographic map). It can be considered as a minimum spanning forest algorithm, and obtained in linear time with respect to the input size [16, 17]. For an image volume V = V (x, y, z, c), we propose to calculate the topographical map T (disaffinity map) as T (x, y, z) = max t∈{x,y,z} max c Gt(x, y, z, c), (1) where x, y, z denote the spatial coordinates, c denotes the color coordinate, and Gx, Gy, Gz denote the spatial gradients of V (nearest neighbor differencing). That is, any edge with significant deviation in any color channel will correspond to a "mountain" in the topographic map. A flooding parameter, f, assigns the local minima of T to catchment basins, which partition V together with the boundary voxels. We assign the boundaries to neighboring basins based on color proximity. The background is the largest and darkest basin. We call the remaining objects supervoxels [18, 19]. Let F denote the binary image identifying all of the foreground voxels. 2 Figure 1: Multiple noise sources affect the color consistency in Brainbow images. a, An 85×121 Brainbow image patch from a single slice (physical size: 8.5µ × 12.1µ). Expression level differs significantly between the membrane and the cytoplasm along a neurite (arrows). b, A maximum intensity projection view of the 3-d image stack. Color shifts along a single neurite, which travels to the top edge and into the page (arrows). c, A 300 × 300 image patch from a single slice of a different Brainbow image (physical size: 30µ × 30µ). d, The intensity variations of the different color channels along the horizontal line in c. e, Same as d for the vertical line in c. f, The image patch in c after denoising. g–h, Same as d and e after denoising. For the plots, the range of individual color channels is [0, 1]. Objects without interior voxels (e.g., single-voxel thick dendritic segments) may not be detected by Eq. 1 (Supp. Fig. 1). We recover such "bridges" using a topology-preserving warping (in this case, only shrinking is used.) of the thresholded image stack into F [20, 21]: B = W(Iθ, F ), (2) where Iθ is binary and obtained by thresholding the intensity image at θ. W returns a binary image B such that B has the same topology as Iθ and agrees with F as much as possible. Each connected component of B ∧ ¯F (foreground of B and background of F ) is added to a neighboring supervoxel based on color proximity, and discarded if no spatial neighbors exist (Supp. Text). We ensure the color homogeneity within supervoxels by dividing non-homogeneous supervoxels (e.g., large color variation across voxels) into connected subcomponents based on color until the desired homogeneity is achieved (Supp. Text). We summarize each supervoxel's color by its mean color. We apply local heuristics and spatio-color constraints iteratively to further reduce the data size and demix overlapping neurons in voxel space (Fig. 2f,g and Supp. Text). Supp. Text provides details on the parallelization and complexity of these steps and the method in general. 3 A C F B E H D G 5 x−position (µm) 10 15 1 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 0 normalized intensity 0 0 2 4 6 8 y−position (µm) 10 12 14 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 normalized intensity 0 0 2 4 6 8 y−position (µm) 10 12 14 0.7 0.6 0.5 0.4 0.3 0.2 0.1 normalized intensity 5 x−position (µm) 10 15 1 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 0 normalized intensity of Figure 2: Best viewed digitally. a, A schematic of the processing steps b, Max. intensity pro- jection of a raw Brainbow image c, Max. intensity projection of the denoised image d, A zoomed-in version the patch indicated by the dashed square in b. e, The corresponding denoised image. f, One-third of the supervoxels in the top-left quadrant (randomly cho- sen). g, Same as f after the merging h1-h4, step. Same as b,c,f,g for simulated data. Scale bars, 20µm. 2.3 Clustering the supervoxel set We consider the supervoxels as the nodes of a graph and express their spatio-color similarities through the existence (and the strength) of the edges connecting them, summarized by a highly sparse adjacency matrix. Removing edges between supervoxels that aren't spatio-color neighbors avoids spurious links. However, this procedure also removes many genuine links due to high color variability (Fig. 1). Moreover, it cannot identify disconnected segments of the same neuron (e.g., due to limited field-of-view). Instead, we adjust the spatio-color neighborhoods based on the "reliability" of the colors of the supervoxels. Let S denote the set of supervoxels in the dataset. We define the sets of reliable and unreliable supervoxels as Sr = {s ∈ S : n(s) > ts, h(s) < td} and Su = S \ Sr, respectively, where n(s) denotes the number of voxels in s, h(s) is a measure of the color heterogeneity (e.g., the maximum difference between intensities across all color channels), ts and td are the corresponding thresholds. We describe a graph G = (V, E), where V denotes the vertex set (supervoxels) and E = Es∪Ec∪E¯s denotes the edges between them: Es = {(ij) : δij < s, i 6= j} Ec = {(ij) : si, sj ∈ Sr, dij < c, i 6= j} E¯s = {(ij), (ji) : si ∈ Su, (ij) /∈ Es, Oi(j) < kmin − Ki, i 6= j}, (3) where δij, dij are the spatial and color distances between si and sj, respectively. s and c are the corresponding maximum distances. An unreliable supervoxel with too few spatial neighbors is allowed to have up to kmin edges via proximity in color space. Here, Oi(j) is the order of supervoxel sj in terms of the color distance from supervoxel si, and Ki is the number of s-spatial neighbors of si. (Note the symmetric formulation in E¯s.) Then, we construct the adjacency matrix as A(i, j) =(cid:26) e−αd2 0, 4 ij , (ij) ∈ E otherwise (4) where α controls the decay in affinity with respect to distance in color. We use k-d tree structures to efficiently retrieve the color neighborhoods [22]. Here, the distance between two supervoxels is minv∈V,u∈U D(v, u), where V and U are the voxel sets of the two supervoxels and D(v, u) is the Euclidean distance between voxels v and u. A classical way of partitioning graph nodes that are nonlinearly separable is by minimizing a function (e.g., the sum or the maximum) of the edge weights that are severed during the partitioning [23]. Here, we use the normalized cuts algorithm [24, 13] with two simple modifications: the k-means step is weighted by the sizes of the supervoxels and initialized by a few iterations of k-means clustering of the supervoxel colors only (Supp. Text). The resulting clusters partition the image stack (together with the background), and represent a segmentation of the individual neurons within the image stack. An estimate of the number of neurons can be obtained from a Dirichlet process mixture model [25]. While this estimate is often rough [26], the segmentation accuracy appears resilient to imperfect estimates (Fig. 4c). 2.4 Simulating Brainbow tissues We create basic simulated Brainbow image stacks from volumetric reconstructions of single neurons (Algorithm 1). For simplicity, we model the neuron color shifts by a Brownian noise component on the tree, and the background intensity by a white Gaussian noise component (Supp. Text). We quantify the segmentation quality of the voxels using the adjusted Rand index (ARI), whose maximum value is 1 (perfect agreement), and expected value is 0 for random clusters [27]. (Supp. Text) background noise variability σ1, neural color variability σ2, r, saturation level M Shift and rotate neuron ni to minimize overlap with existing neurons in the stack Generate a uniformly random color vector vi of length C Identify the connected components of cij of ni within the stack for cij ∈ {cij}j do Algorithm 1 Brainbow image stack simulation Require: number of color channels C, set of neural shapes S = {ni}i, stack (empty, 3d space + color), 1: for ni ∈ S do 2: 3: 4: 5: 6: 7: 8: 9: 10: end for 11: Add white noise to each voxel generated by N (0, σ2 2I) 12: if brightness exceeds M then 13: 14: end if 15: return stack Saturate at M Pre-assign vi to r% of the voxels of cij C-dimensional random walk on cij with steps N (0, σ2 end for Add neuron ni to the stack (with additive colors for shared voxels) 1I) (Supp. Text) 3 Datasets To simulate Brainbow image stacks, we used volumetric single neuron reconstructions of mouse retinal ganglion cells in Algorithm 1. The dataset is obtained from previously published studies [28, 29]. Briefly, the voxel size of the images is 0.4µ × 0.4µ × 0.5µ, and the field of view of individual stacks is 320µ × 320µ × 70µ or larger. We evaluate the effects of different conditions on a central portion of the simulated image stack. Both real datasets are images of the mouse hippocampal tissue. The first dataset has 1020×1020×225 voxels (voxel size: 0.1× 0.1× 0.3µ3), and the tissue was imaged at 4 different frequencies (channels). The second dataset has 1080 × 1280 × 134 voxels with an effective voxel size of 70 × 70 × 40nm, where the tissue was 4× linearly expanded [5], and imaged at 3 different channels. The Brainbow constructs were delivered virally, and approximately 5% of the neurons express a fluorescence gene. 4 Results Parameters used in the experiments are reported in Supp. Text. Fig. 1b, d, and e depict the variability of color within individual neurites in a single slice and through the imaging plane. Together, they demonstrate that the voxel colors of even a small segment of a 5 Figure 3: Segmentation of a simulated Brainbow image stack. Adjusted Rand index of the fore- ground is 0.80. Pseudo- color representation of 4- channel data. Top: max- imum intensity projection of the ground truth. Only the supervoxels that are oc- cupied by a single neuron are shown. Bottom: max- imum intensity projection of the reconstruction. The top-left corners show the whole image stack. All other panels show the max- imum intensity projections of the supervoxels assigned to a single cluster (inferred neuron). Figure 4: Segmentation accuracy of simulated data a, Expression density (ratio of voxels occupied by at least one neuron) vs. ARI. b, σ1 (Algorithm 1) vs. ARI. c, Channel count vs. ARI for a 9-neuron simulation, where K ∈ [6, 12]. ARI is calculated for the foreground voxels. See Supp. Fig. 7 for ARI values for all voxels. neuron's arbor can occupy a significant portion of the dynamic range in color with the state-of-the- art Brainbow data. Fig. 1c-e show that collaborative denoising removes much of this noise while preserving the edges, which is crucial for segmentation. Fig. 2b-e and h demonstrate a similar effect on a larger scale with real and simulated Brainbow images. Fig. 2 shows the raw and denoised versions of the 1020 × 1020 × 225 image, and a randomly chosen subset of its supervoxels (one-third). The original set had 6.2 × 104 supervoxels, and the merging routine decreased this number to 3.9 × 104. The individual supervoxels grew in size while avoiding mergers with supervoxels of different neurons. This set of supervoxels, together with a (sparse) spatial connectivity matrix, characterizes the image stack. Similar reductions are obtained for all the real and simulated datasets. Fig. 3 shows the segmentation of a simulated 200×200×100 (physical size: 80µ×80µ×50µ) image patch. (Supp. Fig. 2 shows all three projections, and Supp. Fig. 3 shows the density plot through the z-axis.) In this particular example, the number of neurons within the image is 9, σ1 = 0.04, σ2 = 0.1, and the simulated tissue is imaged using 4 independent channels. Supp. Fig. 4 shows a patch from a single slice to visualize the amount of noise. The segmentation has an adjusted Rand index of 0.80 when calculated for the detected foreground voxels, and 0.73 when calculated for all voxels. (In some cases, the value based on all voxels is higher.) The ground truth image displays only those supervoxels all of whose voxels belong to a single neuron. The bottom part of Fig. 3 shows 6 0.9 0.85 0.8 0.75 0.7 0.65 0.6 0.55 0.5 3 4 5 true (9) channel count 6 7 8 10 11 12 0.95 0.9 0.85 0.8 0.75 0.7 0.65 0.6 0.55 0 0.02 0.04 0.06 0.08 0.1 3 ch. step size (σ1) −− range per channel: [0, 1] 4 ch. 5 ch. 1 0.95 0.9 0.85 0.8 0.75 0.7 0.65 0.06 0.08 0.1 0.12 0.14 0.16 0.18 expression density (ratio of occupied voxels) 3 ch. adjusted Rand index 4 ch. 5 ch. Figure 5: Segmentation of a Brainbow stack – best viewed digitally. Pseudo-color represen- tation of 4-channel data. The physical size of the stack is 102µ × 102µ × 68µ. The top-left corner shows the maximum intensity projection of the whole image stack, all other panels show the maximum intensity projections of the supervoxels assigned to a single cluster (inferred neuron). 7 that many of these supervoxels are correctly clustered to preserve the connectivity of neuronal arbors. There are two important mistakes in clusters 4 (merger) and 9 (spurious cluster). These are caused by aggressive merging of supervoxels (Supp. Fig. 5), and the segmentation quality improves with the inclusion of an extra imaging channel and more conservative merging (Supp. Fig. 6). We plot the performance of our method under different conditions in Fig. 4 (and Supp. Fig. 7). We set the noise standard deviation to σ1 in the denoiser, and ignored the contribution of σ2. Increasing the number of observation channels improves the segmentation performance. The clustering accuracy degrades gradually with increasing neuron-color noise (σ1) in the reported range (Fig. 4b). The accuracy does not seem to degrade when the cluster count is mildly overestimated, while it decays quickly when the count is underestimated (Fig. 4c). Fig. 5 displays the segmentation of the 1020 × 1020 × 225 image. While some mistakes can be spotted by eye, most of the neurites can be identified and simple tracing tools can be used to obtain final skeletons/segmentations [30, 31]. In particular, the identified clusters exhibit homogeneous colors and dendritic pieces that either form connected components or miss small pieces that do not preclude the use of those tracing tools. Some clusters appear empty while a few others seem to comprise segments from more than one neuron, in line with the simulation image (Fig. 2.4). Supp. Fig. 8 displays the segmentation of the 4× expanded, 1080× 1280× 134 image. While the two real datasets have different characteristics and voxel sizes, we used essentially the same parameters for both of them throughout denoising, supervoxelization, merging, and clustering (Supp. Text). Similar to Fig. 5, many of the processes can be identified easily. On the other hand, Supp. Fig. 8 appears more fragmented, which can be explained by the smaller number of color channels (Fig. 4). 5 Discussion Tagging individual cells with (pseudo)colors stochastically is an important tool in biological sciences. The versatility of genetic tools for tagging synapses or cell types and the large field-of-view of light microscopy positions multispectral labeling as a complementary approach to electron microscopy based, small-scale, dense reconstructions [3]. However, its use in neuroscience has been limited due to various sources of variability in expression. Here, we demonstrate that automated segmentation of neurons in such image stacks is possible. Our approach considers both accuracy and scalability as design goals. The basic simulation proposed here (Algo. 1) captures the key aspects of the problem and may guide the relevant genetics research. Yet, more detailed biophysical simulations represent a valuable direction for future work. Our simulations suggest that the segmentation accuracy increases signifi- cantly with the inclusion of additional color channels, which coincides with ongoing experimental efforts [9–11]. We also note that color constancy of individual neurons plays an important role both in the accuracy of the segmentation (Fig. 4) and the supervoxelized problem size. While we did not focus on post-processing in this paper, basic algorithms (e.g., reassignment of small, isolated supervoxels) may improve both the visualization and the segmentation quality. Similarly, more elaborate formulations of the adjacency relationship between supervoxels can increase the accuracy. Finally, supervised learning of this relationship (when labeled data is present) is a promising direction, and our methods can significantly accelerate the generation of training sets. 6 Acknowledgments The authors thank Suraj Keshri and Min-hwan Oh (Columbia University) for useful conversations. Funding for this research was provided by ARO MURI W911NF-12-1-0594, DARPA N66001- 15-C-4032 (SIMPLEX), and a Google Faculty Research award; in addition, this work was supported by the Intelligence Advanced Research Projects Activity (IARPA) via Department of Interior/ Interior Business Center (DoI/IBC) contract number D16PC00008. The U.S. Government is authorized to reproduce and distribute reprints for Governmental purposes notwithstanding any copyright annotation thereon. Disclaimer: The views and conclusions contained herein are those of the authors and should not be interpreted as necessarily representing the official policies or endorsements, either expressed or implied, of IARPA, DoI/IBC, or the U.S. Government. 8 References [1] J Livet et al. Transgenic strategies for combinatorial expression of fluorescent proteins in the nervous system. Nature, 450(7166):56–62, 2007. [2] A H Marblestone et al. Rosetta brains: A strategy for molecularly-annotated connectomics. arXiv preprint arXiv:1404.5103, 2014. [3] Shin-ya Takemura, Arjun Bharioke, Zhiyuan Lu, Aljoscha Nern, Shiv Vitaladevuni, Patricia K Rivlin, William T Katz, Donald J Olbris, Stephen M Plaza, Philip Winston, et al. A visual motion detection circuit suggested by drosophila connectomics. Nature, 500(7461):175–181, 2013. [4] D Cai et al. Improved tools for the brainbow toolbox. Nature methods, 10(6):540–547, 2013. [5] F Chen et al. Expansion microscopy. Science, 347(6221):543–548, 2015. [6] K Chung and K Deisseroth. Clarity for mapping the nervous system. Nat. methods, 10(6):508–513, 2013. [7] E Betzig et al. Imaging intracellular fluorescent proteins at nanometer resolution. Science, 313(5793):1642– 1645, 2006. [8] M J Rust et al. Sub-diffraction-limit imaging by stochastic optical reconstruction microscopy (storm). Nature methods, 3(10):793–796, 2006. [9] A M Zador et al. Sequencing the connectome. PLoS Biol, 10(10):e1001411, 2012. [10] J H Lee et al. Highly multiplexed subcellular rna sequencing in situ. Science, 343(6177):1360–1363, 2014. [11] K H Chen et al. Science, Spatially resolved, highly multiplexed rna profiling in single cells. 348(6233):aaa6090, 2015. [12] M Maggioni et al. Nonlocal transform-domain filter for volumetric data denoising and reconstruction. Image Processing, IEEE Transactions on, 22(1):119–133, 2013. [13] U Von Luxburg. A tutorial on spectral clustering. Statistics and computing, 17(4):395–416, 2007. [14] D Lee and H S Seung. Algorithms for non-negative matrix factorization. In Advances in neural information processing systems, pages 556–562, 2001. [15] F Meyer. Topographic distance and watershed lines. Signal processing, 38(1):113–125, 1994. [16] F Meyer. Minimum spanning forests for morphological segmentation. In Mathematical morphology and its applications to image processing, pages 77–84. Springer, 1994. [17] J Cousty et al. Watershed cuts: Minimum spanning forests and the drop of water principle. Pattern Analysis and Machine Intelligence, IEEE Transactions on, 31(8):1362–1374, 2009. [18] Xiaofeng Ren and Jitendra Malik. Learning a classification model for segmentation. In Computer Vision, 2003. Proceedings. Ninth IEEE International Conference on, pages 10–17. IEEE, 2003. [19] J S Kim et al. Space-time wiring specificity supports direction selectivity in the retina. Nature, 509(7500):331–336, 2014. [20] Gilles Bertrand and Grégoire Malandain. A new characterization of three-dimensional simple points. Pattern Recognition Letters, 15(2):169–175, 1994. [21] Viren Jain, Benjamin Bollmann, Mark Richardson, Daniel R Berger, Moritz N Helmstaedter, Kevin L Briggman, Winfried Denk, Jared B Bowden, John M Mendenhall, Wickliffe C Abraham, et al. Boundary learning by optimization with topological constraints. In Computer Vision and Pattern Recognition (CVPR), 2010 IEEE Conference on, pages 2488–2495. IEEE, 2010. [22] J L Bentley. Multidimensional binary search trees used for associative searching. Communications of the ACM, 18(9):509–517, 1975. [23] Z Wu and R Leahy. An optimal graph theoretic approach to data clustering: Theory and its application to image segmentation. IEEE Trans. Pattern Anal. Mach. Intell., 15(11):1101–1113, 1993. [24] A Y Ng et al. On spectral clustering: Analysis and an algorithm. Advances in neural information processing systems, 2:849–856, 2002. [25] K Kurihara et al. Collapsed variational dirichlet process mixture models. In IJCAI, volume 7, pages 2796–2801, 2007. [26] Jeffrey W Miller and Matthew T Harrison. A simple example of dirichlet process mixture inconsistency for the number of components. In Advances in neural information processing systems, pages 199–206, 2013. [27] Lawrence Hubert and Phipps Arabie. Comparing partitions. Journal of classification, 2(1):193–218, 1985. [28] U Sümbül et al. A genetic and computational approach to structurally classify neuronal types. Nature communications, 5, 2014. [29] U Sümbül et al. Automated computation of arbor densities: a step toward identifying neuronal cell types. Frontiers in neuroscience, 2014. [30] M H Longair et al. Simple neurite tracer: open source software for reconstruction, visualization and analysis of neuronal processes. Bioinformatics, 27(17):2453–2454, 2011. [31] H Peng et al. V3d enables real-time 3d visualization and quantitative analysis of large-scale biological image data sets. Nature biotechnology, 28(4):348–353, 2010. 9 Supplementary information: Automated scalable segmentation of neurons from multispectral images Supplementary Text Parameter choices and feature representations Similar results are obtained around a neighborhood of these suggested values. Unless otherwise noted, the same parameter values were used for all three reported experiments: two real hippocampal datasets acquired by different labs and under different conditions, and a set of simulated retinal datasets with different parameters. The subsection of the main text that refers to these parameters are indicated in square brackets. Color features: For every color triplet, we obtain the L-u-v representation (Schanda, 2007), and calculate the top C principal components of the concatenated L-u-v representations, where C is the number of color channels. If the data has more than 3 channels, for affinity calculations between neighboring supervoxels, we normalize the colors before the L-u-v transformation. [Dimensionality reduction] Supervoxel reliability: ts = 50, td = 0.5 (before L-u-v transformation) [Clustering the supervoxel set] Edge set parameters: s = √3 (26-neighborhood for isotropic data), c = 20 ×pC/4 (by inspect- ing typical color radius within individual neurons and manual adjustment), kmin = 5. [Clustering the supervoxel set] Edge strength decay: α = 2 × 10−3 (by inspecting typical color radius and manual adjustment) [Clustering the supervoxel set] Flooding parameter for watershed: f = 0.01 with 26-neighborhood. (This affects computation time more than quality because subdividing via the maximum color perimeter can catch inhomoge- neous supervoxels.) [Dimensionality reduction] Maximum color perimeter for supervoxel homogeneity: (Supp. Algo. 1) p = 0.5 for each channel when the intensity is in [0, 1] (by inspecting data – see Fig. 1). [Dimensionality reduction] Image thresholding for warping: θ = 0.1 ×pC/4 before L-u-v transformation (for the expansion microscopy data, θ = 0.2) [Dimensionality reduction] Noise standard deviation for denoising: σ = 1/8 when the intensity is in [0, 1] for individual channels. [Denoising the image stack] Cluster (neuron) counts: For the dataset in Fig. 5, the mixture model (Kurihara et al., 2007) suggested 52 clusters based on the colors of the supervoxels. The same routine returned 29 clusters when run on 1 5 of the supervoxels. We chose K = 34 for a compact presentation. For the dataset in 30th Conference on Neural Information Processing Systems (NIPS 2016), Barcelona, Spain. arXiv:1611.00388v2 [q-bio.NC] 22 Jan 2017 Supp. Fig. 8, we used K = 19, which is what the mixture model suggested based on the colors of the supervoxels. [Clustering the supervoxel set] Spatial distance calculation The spatial distance between two supervoxels is calculated as minv∈V1,u∈V2 D(v, u), where V1 and V2 are the voxel sets of the two supervoxels and D(v, u) is the Euclidean distance between voxels v and u. Only the boundary voxels need to be considered, and extremal values in each coordinate are used to identify many supervoxel pairs farther than s without exact calculation over the voxels. Only the spatial distances between nearby supervoxels need to be computed. Color-based subdivision of supervoxels Let the n × C matrix Vi denote the colors of all n voxels of the supervoxel si. Supp. Algo. 1 divides the supervoxels into smaller supervoxels until the desired homogeneity is achieved. Supplementary Algorithm 1 Subdivide supervoxels Require: S = {si}i (set of supervoxels), {Vi}i, pmax (threshold) 1: Snew = {} 2: for si ∈ S do 3: 4: 5: 6: 7: p = maxc∈C max(Vi(:, c)) − min(Vi(:, c)) if p < pmax then Add si to Snew else Divide the voxels into 2 sets T1 and T2 based on color (e.g., using k-means, hierarchical clustering, etc.) Add the connected components of T1 and T2 to S end if Remove si from S 8: 9: 10: 11: end for 12: return Snew Simulation data RGC arbors stratify in the retina, distributing their dendritic length within a slab. To achieve denser simulations, we did not shift the neurons much in z (The density numbers calculated in Fig. 4 are obtained by considering the [35µm, 50µm] region in Supp. Fig. 3.) We obtain simulated stacks by varying the expression density (S ∈ {5, 9, 13}), the channel count (C ∈ {3, 4, 5}), the neuron color consistency (σ1 ∈ {0.01, 0.02, 0.03, 0.04, 0.1}), and the background noise (σ2 ∈ {0.05, 0.1}). The random walk on a connected component assigns a color to a voxel by (i) calculating the mean color of the neighboring voxels that were previously visited, and (ii) adding the random noise step to this mean value. Adjusted Rand index We quantify the segmentation quality of the voxels of the simulated dataset using the adjusted Rand index. The Rand index is a measure of the element pairs on which two partitions P and P of the same set with N elements agree: R(P, P ) = 1 −(cid:0)N 2(cid:1)Pi<j δ(li, lj) − δ(li, lj), where li (li) denotes the label of element i according to P ( P ), and δ is the indicator function. (δ(li, lj) = 1 if li = lj, δ(li, lj) = 0 otherwise.) The adjusted Rand index corrects for chance, has a more sensitive dynamic 2 range, and is defined as A = (R − E)/(M − E), where E is the expected value of the index and M is the maximum value of the index, based on the number of elements in individual segments. Its maximum value is 1 (perfect agreement), and the expected value of the index is 0 for random clusters. For the foreground based calculation, only the voxels that are assigned to the foreground after the watershed transform and warping are considered. For the image based calculation, all voxels are considered and the background is treated as a separate object. Merging supervoxels We apply local heuristics and spatio-color constraints iteratively to further reduce the data size and demix overlapping neurons in voxel space (Fig. 2): (i) supervoxels occupied by more than one neuron are detected and demixed by monitoring the improvement in non-negative least squares fit quality. (Supp. Algo 2.) (ii) neighboring supervoxels with similar colors and orientations, supervoxels with single spatial neighbors, and supervoxels all of whose neighbors have similar colors are merged. (iii) supervoxels that are spatial neighbors and that are assigned to the same cluster by an overclustering color k-means routine are merged. We implement (iii) to run in parallel over subgraphs of the full graph for scalability. A rough estimate of the number of neurons required by the oversegmentation routine is obtained by a Dirichlet process mixture model (Kurihara et al., 2007). The k-means algorithm uses a multiple of this rough estimate. Note that only a rough estimate (Miller and Harrison, 2013) is needed because of oversegmentation (Supp Algo. 3). This algorithm can be implemented to run in parallel over subgraphs of the full dataset. if si has less than M voxels then 2 subject to x ≥ 0 size), ∆ (maximum color distance), f (improvement factor) Initialize r = V (i, :)2 for each neighbor pair (i1, i2) do 2, P = (0, 0) t = minx V ([i1, i2], :)x − V (i, :)2 if t < r then if the minimum color distance between si and its neighbors is larger than ∆ then Retain the neighbors that have neighbors with color distance less than ∆ if si has more than one spatial neighbors then Supplementary Algorithm 2 Demixing of supervoxels Require: S = {si}i, V (matrix of normalized supervoxel colors), A (spatial affinity matrix), M (maximum 1: for si ∈ S do 2: 3: 4: 5: 6: 7: 8: 9: 10: 11: 12: 13: 14: 15: 16: 17: 18: 19: 20: 21: end for 22: return S, V , A Assign the voxels of si to both of sP (1) and sP (2) Update the spatial affinities of sP (1) and sP (2) in A accordingly Remove si from S, V , and A r = t, P = (i1, i2) end if end for if r < (∆/f )2 then end if end if end if end if 3 factor) Supplementary Algorithm 3 Spatio-color merging of supervoxels Require: S = {si}i, (set of supervoxels) K (rough estimate of the number of clusters), k (oversegmentation 1: Snew = {} 2: Divide S into kK clusters based on the colors of the supervoxels, using k-means 3: for κ1 ∈ {1, . . . , kK} do 4: 5: 6: end for 7: return Snew Find the connected components within the cluster κ1 Merge the supervoxels within the connected components of that cluster, and add to Snew Supplementary Figure 1: Top: Maximum intensity projection of a raw Brainbow image. Bottom left: Fore- ground after watershed transform. Arrows point to six different thin dendritic pieces ("bridges") that were missed. Bottom right: Foreground after warping correction. Scale bar, 30µm 4 Supplementary Figure 2: The z (top left), x (top right), and y (bottom left) maximum intensity projections of the raw simulation image shown in Fig. 3. Adjusted Rand index of the segmentation is 0.80. 5 Supplementary Figure 3: The z-profile of the ground truth simulation image with 13 neurons, showing that a region is preferentially occupied. The range [0µm, 50µm] corresponds to slices 1 to 100 so that most of the neuronal arbors are between slices 70 and 100. 6 10 40 distance from the top of the stack (µ m) 20 30 50 0.35 0.3 0.25 0.2 0.15 0.1 0.05 ratio of occupied voxels 0 0 Supplementary Figure 4: A 60 × 60 patch from a single slice (slice 90) of the simulation image shown in Fig. 3. Top: raw. Bottom: denoised. Physical size: 15µm × 15µm 7 A B Supplementary Figure 5: Aggressive merging generates supervoxels with inconsistent colors. Top: Cluster 4 in the bottom part of Fig. 3 of the main text. Bottom: Cluster 9 in the bottom part of Fig. 3 of the main text. Close inspection reveals that some of the multi-colored regions comprise single supervoxels. 8 Supplementary Figure 6: Segmentation of a simulated Brainbow image stack. Adjusted Rand index of the foreground is 0.87. Pseudo-color representation of 5-channel data with more conservative supervoxel merging compared to Fig. 3 of the main text. Maximum intensity projection of the segmentation. The top-left corner shows the whole image stack. All other panels show the maximum intensity projections of the supervoxels assigned to a single cluster (inferred neuron). Supplementary Figure 7: Segmentation accuracy of simulated data a, Expression density (ratio of voxels occupied by at least one neuron) vs. ARI. b, σ1 vs. ARI. c, Channel count vs. ARI for a 9-neuron simulation, where K ∈ [6, 12]. ARI is calculated for all voxels. 9 0.74 0.735 0.73 0.725 0.72 0.715 0.71 0.705 0.7 0.695 3 4 true (9) channel count 6 7 8 10 11 12 0.75 0.74 0.73 0.72 0.71 0.7 0.69 0.68 0.67 0.66 0.06 0.08 0.1 0.12 0.14 0.16 0.18 3 ch. expression density (ratio of occupied voxels) adjusted Rand index 4 ch. 5 ch. 0.8 0.78 0.76 0.74 0.72 0.7 0.68 0.66 0.64 0.62 0 0.02 0.04 0.06 0.08 0.1 3 ch. step size (σ1) −− range per channel: [0, 1] 4 ch. 5 ch. Supplementary Figure 8: Segmentation of a 4× linearly expanded Brainbow stack – best viewed digitally. The physical size of the stack is 90µ × 76µ × 5µ. The top-left corner shows the maximum intensity projection of the whole image stack, all other panels show the maximum intensity projections of the supervoxels assigned to a single cluster (inferred neuron). Parallelization and complexity The denoising step (collaborative filtering) parallelizes over substacks (voxels) because the extra dimension is formed by local patches. The watershed algorithm can also be parallelized over substacks. Moreover, it has a linear run-time with respect to the input size (Main Text). Similarly, warping (shrinking) can be applied on those substacks. Querying the boundary voxels for flipping at each stage, and ordering them by brightness result in a fast implementation. It has a linear run-time with respect to the input size. (Simple voxel query can be performed over 3 × 3 × 3 patches – See references in main text.) 10 Supervoxel merging can be performed in parallel on individual substacks. Merging subroutines use local rules to make local merges except for the spatio-color merging step, which uses the k-means algorithm. Similarity calculations require extracting spatial and color neighborhoods of the supervoxels. Spatial distance calculation is discussed above. These value are precalculated. As mentioned in the main text, k-d tree structures are used to retrieve the color neighborhoods efficiently. Finally, clustering is performed by the normalized cuts algorithm. We use an implicitly restarted block Lanczos method for computing the first few eigenvectors (Baglama et al., 2003). References Baglama, J., Calvetti, D., and Reichel, L. (2003). Algorithm 827: irbleigs: A matlab program for computing a few eigenpairs of a large sparse hermitian matrix. ACM Transactions on Mathematical Software (TOMS), 29(3):337–348. Kurihara, K., Welling, M., and Teh, Y. W. (2007). Collapsed variational dirichlet process mixture models. In IJCAI, volume 7, pages 2796–2801. Miller, J. W. and Harrison, M. T. (2013). A simple example of dirichlet process mixture inconsistency for the number of components. In Advances in neural information processing systems, pages 199–206. Schanda, J. (2007). Colorimetry: understanding the CIE system. John Wiley & Sons. 11
1210.3555
2
1210
2013-10-30T15:30:56
Task-Based Core-Periphery Organisation of Human Brain Dynamics
[ "q-bio.NC", "cond-mat.dis-nn", "nlin.AO", "stat.AP" ]
As a person learns a new skill, distinct synapses, brain regions, and circuits are engaged and change over time. In this paper, we develop methods to examine patterns of correlated activity across a large set of brain regions. Our goal is to identify properties that enable robust learning of a motor skill. We measure brain activity during motor sequencing and characterize network properties based on coherent activity between brain regions. Using recently developed algorithms to detect time-evolving communities, we find that the complex reconfiguration patterns of the brain's putative functional modules that control learning can be described parsimoniously by the combined presence of a relatively stiff temporal core that is composed primarily of sensorimotor and visual regions whose connectivity changes little in time and a flexible temporal periphery that is composed primarily of multimodal association regions whose connectivity changes frequently. The separation between temporal core and periphery changes over the course of training and, importantly, is a good predictor of individual differences in learning success. The core of dynamically stiff regions exhibits dense connectivity, which is consistent with notions of core-periphery organization established previously in social networks. Our results demonstrate that core-periphery organization provides an insightful way to understand how putative functional modules are linked. This, in turn, enables the prediction of fundamental human capacities, including the production of complex goal-directed behavior.
q-bio.NC
q-bio
Task-Based Core-Periphery Organization of Human Brain Dynamics Danielle S. Bassett1,2,∗, Nicholas F. Wymbs5, M. Puck Rombach3,4, Mason A. Porter3,4, Peter J. Mucha6,7, Scott T. Grafton5 1 Department of Physics, University of California, Santa Barbara, CA 93106, USA 2 Sage Center for the Study of the Mind, University of California, Santa Barbara, CA 93106, USA 3 Oxford Centre for Industrial and Applied Mathematics, Mathematical Institute, University of Oxford, Oxford OX1 3LB, UK 4 CABDyN Complexity Centre, University of Oxford, Oxford, OX1 1HP, UK 5 Department of Psychological and Brain Sciences and UCSB Brain Imaging Center, University of California, Santa Barbara, CA 93106, USA 6 Carolina Center for Interdisciplinary Applied Mathematics, Department of Mathematics, University of North Carolina, Chapel Hill, NC 27599, USA 7 Institute for Advanced Materials, Nanoscience & Technology, University of North Carolina, Chapel Hill, NC 27599, USA ∗ E-mail: Corresponding [email protected] Abstract As a person learns a new skill, distinct synapses, brain regions, and circuits are engaged and change over time. In this paper, we develop methods to examine patterns of correlated activity across a large set of brain regions. Our goal is to identify properties that enable robust learning of a motor skill. We measure brain activity during motor sequencing and characterize network properties based on coherent activity between brain regions. Using recently developed algorithms to detect time-evolving communities, we find that the complex reconfiguration patterns of the brain's putative functional modules that control learning can be described parsimoniously by the combined presence of a relatively stiff temporal core that is composed primarily of sensorimotor and visual regions whose connectivity changes little in time and a flexible temporal periphery that is composed primarily of multimodal association regions whose connectivity changes frequently. The separation between temporal core and periphery changes over the course of training and, importantly, is a good predictor of individual differences in learning success. The core of dynamically stiff regions exhibits dense connectivity, which is consistent with notions of core-periphery organization established previously in social networks. Our results demonstrate that core- periphery organization provides an insightful way to understand how putative functional modules are linked. This, in turn, enables the prediction of fundamental human capacities, including the production of complex goal-directed behavior. Author Summary When someone learns a new skill, his/her brain dynamically alters individual synapses, regional activity, and larger-scale circuits. In this paper, we capture some of these dynamics by measuring and character- izing patterns of coherent brain activity during the learning of a motor skill. We extract time-evolving communities from these patterns and find that a temporal core that is composed primarily of primary sen- sorimotor and visual regions reconfigures little over time, whereas a periphery that is composed primarily of multimodal association regions reconfigures frequently. The core consists of densely connected nodes, and the periphery consists of sparsely connected nodes. Individual participants with a larger separation between core and periphery learn better in subsequent training sessions than individuals with a smaller separation. Conceptually, core-periphery organization provides a framework in which to understand how arXiv:1210.3555v2 [q-bio.NC] 30 Oct 2013 2 putative functional modules are linked. This, in turn, enables the prediction of fundamental human capacities, including the production of complex goal-directed behavior. Introduction Cohesive structures have long been thought to play an important role in information processing in the human brain [1]. At the small scale of individual neurons, temporally coherent activity supports in- formation transfer between cells [2]. At a much larger scale, simultaneously active cortical areas form functional systems that enable behavior [1]. However, the question of precisely what type of cohesive organization is present between the constituents of brain systems -- especially at larger scales -- has been steeped in controversy [3, 4]. Although low-frequency interactions between pairs of brain areas are easy to measure, the simultaneous characterization of dynamic interactions across the entire human brain remained challenging until recent applications of network theory to neuroimaging data [5]. These efforts have led to enormous insights, including the establishment of relationships between stationary functional brain network configuration and intelligence [6] as well as relationships between altered brain network organization and disease [7]. In this paper, we extend this approach to a non-stationary situation: the change of network activity across the brain as a new skill is acquired. Acquisition of new motor skills alters brain activity across spatial scales. At the level of individual neurons, this induces changes in firing behavior in the motor cortex [8]. At the level of large-scale areas, this induces changes in the interactions between primary motor cortex and premotor areas, and these changes can influence the amount of learning [9]. Previous studies have demonstrated that pairwise interactions between some of these premotor regions, as measured by the magnitude of coherence between low-frequency blood-oxygen-level-dependent (BOLD) signals, strengthen with practice [10]. Furthermore, complex contributions by non-motor systems such as prefrontal cortex are involved in the strategic control of behavior during learning [11]. These findings reveal some of the changes in local circuits that occur with learning. However, there remains no global assessment of changes in brain networks as a result of learning. In this paper, we seek to find cohesive structures in global brain networks that capture dynamics that are particularly relevant for characterizing skill learning that takes place over the relatively long time scales of minutes to hours of practice. To address these issues, we extract a set of functional networks from task-based functional Magnetic Resonance Imaging (fMRI) time series that describe functional connectivity between brain regions. We probe the dynamics of these putative interactions by subdividing time series into discrete time intervals (of approximately two minutes in duration; see Fig. 1A) during the acquisition of a simple motor skill. Subjects learned a set of 10-element motor sequences similar to piano arpeggios by practicing for at least 30 days during a 6-week period. The depth of training was manipulated so that 2 sequences were extensively practiced (EXT), 2 sequences were moderately practiced (MOD), and 2 sequences were minimally practiced (MIN) on each day. In addition, subjects performed blocks of all of the sequences during fMRI scanning on approximately days 1, 14, 28, and 42 of practice. Using the fMRI time series, we extract functional networks representing the coherence between 112 cortical and subcortical areas for each sequence block. To characterize brain dynamics, we represent sets of functional networks as multilayer brain networks and we identify putative functional modules -- i.e., groups of brain regions that exhibit similar BOLD time courses -- in each 2 -- 3 minute time window. Such cohesive groups of nodes are called "communities" in the network-science literature [12, 13], and they suggest that different sets of brain regions might be related to one another functionally either through direct anatomical connections or through indirect activation by an external stimulus. A community of brain regions might code for a different function (e.g., visual processing, motor performance, or cognitive control), or it might engage in the same function using a distinct processing stream. Characterizing changes in community structure thus makes it possible to map meaningful dynamic patterns of functional connectivity that relate to changes in cognitive function (e.g., 3 learning). We employ computational tools for dynamic community detection [14, 15] for multilayer represen- tations of temporal networks [16] and summarize our findings using diagnostics that quantify three properties of community structure. (See Materials and Methods for their definitions and Ref. [17] for evidence supporting the utility of these diagnostics in capturing changes in brain dynamics over 3 days of learning.) To measure the strength of functional modularization in the brain and quantify the extent of compartmentalization of putative functional modules, we maximize a quality function called multilayer modularity Q to obtain a partition of the brain into communities. (The associated maximum value of Q is known as the maximum modularity.) A high value of Q indicates that the pattern of functional connectivity in the brain can be clustered sensibly into distinct communities of brain regions that exhibit similar time courses. We also compute the number n of communities (i.e., putative functional modules) in partitions of the multilayer networks. A large value of n indicates that there are a large number of distinct temporal profiles in BOLD activations in the brain. To measure the temporal variability of community structure, we compute the flexibility fi of each region i, as this quantifies the frequency that a brain region i changes its allegiances to network communities over time. A high value of flexibility indicates that a region often changes community affiliation. Our results demonstrate that the temporal evolution of community structure is modulated strongly by the depth of training (as reflected in the total number of practiced trials). We also show that the temporal variability of module allegiance varies across brain regions. Sensorimotor and visual cortices form the bulk of a relatively stiff temporal core in which module affiliations change little over a scanning session, whereas multimodal association areas form the bulk of a relatively flexible temporal periphery in which module affiliations change frequently. The separation between the temporal core and temporal periphery predicts individual differences in extended learning. We combine these methods for identifying a temporal core and periphery with a notion of core-periphery organization that originated in the social sciences [18] to show that the organizational structure of functional networks in 2 -- 3 minute time windows correlates with the organizational structure of the brain's temporal evolution: densely connected regions in individual time windows tend to exhibit little change in module allegiance over time, whereas weakly connected regions tend to exhibit significant changes. Taken together, our results suggest that core- periphery organization is a critical property that is as important as modularity for understanding and predicting cognition and behavior (see Fig. 1B). Results Dynamic Community Structure Changes with Learning Community structure changes with the number of trials practiced, independent of when the practice occurred in the 6 weeks. In Fig. 2, we show multilayer modularity (Q, a measure of the quality of a partition into communities), the number of communities, and mean flexibility (F = 1 i=1 fi, a measure of the temporal variability in module allegiance) as a function of the number of training trials completed after a scanning session. See Materials and Methods for the definitions. After an initial increase from 50 to 200 trials practiced, multilayer modularity decreases with an increase in the number of trials practiced, suggesting that community structure in functional brain networks becomes less pronounced with learning. Both the number of communities and the flexibility of community structure increase with the number of trials practiced, which is consistent with an increased specificity of functional connectivity patterns with extended learning. NPN 4 Temporal Core-Periphery Organization Regional Variation in Flexibility. The mean flexibility over participants varied over brain regions. It ranged from approximately 0.04 to approximately 0.14, which implies that brain regions changed their modular affiliation between 4% and 14% of trial blocks on average (see Fig. 3A). The distribution of flexibility across brain regions is decidedly non-Gaussian: the majority of brain regions have relatively high flexibilities, but there is a left-heavy tail of regions (including a small peak) with low values of flexibility. We characterized the distribution of flexibility over brain regions by calculating the third (skewness) and fourth (kurtosis) central moments. The skewness was 0.50 ± 0.26, and the kurtosis was 3.04 ± 0.57. To interpret these findings, we note that a distribution's skewness is a measure of its asymmetry, and the positive values that we observe indicate that the distributions from all participants are skewed to the right. The kurtosis of a distribution is a combined measure of its peakedness [19] and its bimodality [20], and it is sometimes construed as a measure of the extent that a distribution is prone to outliers. The kurtosis values that we observe vary between 2.5 and 5, which includes the value of 3 that occurs for a Gaussian distribution. Defining the Temporal Core and Temporal Periphery. To determine the significance of a brain region's variation in flexibility, we compared the flexibility of brain regions in the empirical multilayer network to that expected in a nodal null model. We can define a temporal core, bulk, and periphery (see Fig. 3). The core is the set of regions whose flexibility is significantly less than expected in the null model; the periphery is the set of regions whose flexibility is significantly greater than expected in the null model; and the bulk consists of all remaining regions. As discussed in the Text S1, the delineation of the brain into these three groups is robust both to the intensity of training (MIN, MOD, or EXT) and to the duration of training (sessions 1 -- 4). Furthermore, the temporal core, bulk, and periphery tend to form their own communities, although the relationship between core-periphery organization and modular organization appears to be altered by learning (see the Text S1). We show the anatomical locations of the temporal core, temporal bulk, and temporal periphery in Fig. 3. The relatively stiff core is composed of 19 regions located predominantly in primary sensorimotor areas in both left and right hemispheres. Most of the motor-related regions in the core were left-lateralized, which is consistent with the participants' use of their right hand to perform the motor sequence. The more flexible periphery is composed of 25 regions located predominantly in multimodal areas -- including inferior parietal, intraparietal sulcus, temporal parietal junction, inferotemporal, fusiform gyrus, and visual association areas. The bulk contains the remaining 68 cortical and subcortical regions -- including large swaths of frontal, temporal, and parietal cortices. See Table S3 for a complete list of the affiliation of each brain region to the temporal core, bulk, and periphery. The separation of a temporally stiff core of predominantly unimodal regions that process information from single sensory modality (e.g., vision, audition, etc.) and a flexible periphery of predominantly multimodal cortices that process information from multiple modalities is consistent with existing understanding of the association of multimodal cortex with the binding of different types of information and the performance of a broad range of cognitive functions [21]. Temporal Core-Periphery Organization and Learning. One can interpret the anatomical location of the temporally stiff core that consists primarily of unimodal regions and a flexible periphery that consists primarily of multimodal cortices in the context of the known roles of these cortices in similar tasks. The ability to retrieve and rapidly execute complex motor sequences requires extensive practice. These well-learned sequences are known to be generated by "core" areas [8, 22 -- 24]. However, when first learning a sequence, people can use a variety of cognitive strategies that are supported by other brain systems (some of which are located in the periphery to augment performance) [25,26]. In some cases, these strategies are detrimental to skill retention [27]. Consequently, we hypothesized that individuals whose core and periphery are distinct -- indicating a strong separability of visuomotor and cognitive regions -- would learn better than those whose core and periphery were less distinguishable from one another. 5 .= To test this hypothesis, we calculated the Spearman rank coefficient ρ between the skewness and kurtosis of the flexibility distribution estimated from the fMRI data of the first scanning session and the learning parameter κ estimated over the next 10 days of home training (see Materials and Methods). The kurtosis (in essence) measures the separation between the temporal core and the temporal periphery and is negatively correlated with κ (the correlation is ρ 0.027), indicating that individuals with a narrower separation between temporal core and temporal periphery learn better in the subsequent 10 home training sessions than individuals with a greater separation between temporal core and temporal periphery. Skewness (in essence) measures the presence of -- rather than the separation between -- the temporal core and the temporal periphery and is also negatively correlated with learning (ρ 0.034), indicating that individuals whose flexibility was more skewed over brain regions learned better than those whose flexibility over brain regions was less skewed. This finding implies that individuals with "stronger" temporal core-periphery structure (i.e., larger values of skewness) learn better than those with "weaker" temporal core-periphery structure (i.e., smaller values of skewness). −0.498, and the p-value is p .= .= −0.480 and p .= Importantly, the temporal separation of the data from the scanning session (which we used to esti- mate brain flexibility) and the home training (which we used to estimate learning) ensures that these correlations are predictive. Together, these results indicate that individuals with a stronger temporal core and temporal periphery but a smooth transition between them seem to learn better than individuals with a weaker temporal core and temporal periphery but a sharper transition between them. These results suggest that successful brain function might depend on a delicate balance between a set of core regions whose allegiance to putative functional modules changes little over time and a set of peripheral regions whose allegiance to putative functional modules is flexible through time (and also on the smoothness of the transition between these two types of regions). Geometrical Core-Periphery Organization Given our demonstration that there exists a temporal core in dynamic brain networks, it is important to ask what role such core regions might play in individual network layers of the multilayer network [17]. While the roles of nodes in a static network can be studied in multiple ways [28,29], we focus on describing the geometrical core-periphery organization -- which can be used to help characterize the organization of edge strengths throughout a network -- to compare it with the temporal core-periphery organization discussed above. The geometrical core of a network is composed of a set of regions that are strongly and mutually interconnected. Measures of network centrality can be useful for identifying nodes in a geometrical core because such measures help capture a node's relative importance within a network in terms of its immediate connections, its distance to other nodes in the graph, or its influence on other nodes in the graph [30, 31]. Drawing on studies of social networks [18], we examine geometrical core-periphery organization in networks extracted from individual time windows by testing whether core nodes are densely connected to one another and whether peripheral nodes are sparsely connected to one another. Rather than proposing a strict separation between a single core and single periphery, we assess the role of a node along a core-periphery spectrum using a centrality measure known as the (geometrical) core score C, which was introduced in Ref. [30]. Network nodes with high C values are densely connected to one another, whereas nodes with low C values are sparsely connected to one another. The method in Ref. [30] uses a two- parameter function to interpolate between core nodes and peripheral nodes. One parameter (which is denoted by α) sets the sharpness of the boundary between the geometrical core and the geometrical periphery. Small values of α indicate a fuzzy boundary, whereas large values indicate a sharp transition. The second parameter (which is denoted by β) sets the size of the geometrical core. Smaller values of β correspond to smaller cores. We can quantify the fit of the transition function that defines the set of core scores to the data using a summary diagnostic that is called the R-score (see Materials and Methods for definitions). Large values of R indicate a good fit and therefore provide confidence that one has uncovered a good estimate of a network's core-periphery organization. 6 .= 0.94) to assign a core score to each node. In Fig. 4A, we show a typical R-score landscape in the (α, β) parameter plane. This landscape favors a relatively small core and a medium value of the transition-sharpness parameter. To choose sensible values of α and β for studying core-periphery organization, we examine the distributions of the relative frequencies of α and β values that maximize the R-score for each network layer, participant, scanning session, and sequence type (see Fig. 4B). We use the mean values of these distributions (α 0.40 and β In Fig. 4C, we show the shape of the "mean core" that we obtain using these parameter values. This figure demonstrates that the typical (geometrical) core-periphery organization in the networks under study is a mixture between a discrete core-periphery organization, in which every node is either in the core or in the periphery, and a continuous core-periphery organization, in which there is a continuous spectrum to describe how strongly nodes belong to a core. In these networks (which usually possess a single core), the majority of nodes do not belong to the core, but those nodes that do (roughly 10% of the nodes) have a continuum level of association strengths with the core. .= .= In some cases, we identified multiple competing cores, which we found by using simulated annealing to explore local maxima of the R-score rather than only identifying a global maximum. Because of this stochasticity in the methodology for examining core-periphery organization, we performed computations with the chosen parameter values (α 0.94) 10 times and used the solution with the highest R-score out of these 10 iterations for each network layer, subject, scanning session, and sequence type. 0.40 and β .= An interesting question is whether geometrical core-periphery organization remains relatively constant throughout time or whether the organization changes with learning. We observed that regions that have a high geometrical core score in the first scanning session and in EXT blocks were likely to have high geometrical core scores in later scanning sessions and in MOD and MIN blocks. (See the Text S1 for supporting results on the reliability of geometrical core-periphery organization.) In light of this consistency, we calculate a mean geometrical core score for each node by taking the mean over all blocks in a given scanning session (1, 2, 3, and 4) and sequence type (EXT, MOD, and MIN). The variance of the mean geometrical core score over nodes in a network then gives an indication of the separation between the mean core and periphery. As we show in Fig. 5, we find that the variance of the mean geometrical core score over trials decreases as a function of learning. A high variance of the mean core scores over nodes indicates a greater separation between the mean core and periphery as well as a high consistency of the core score of each node over trial blocks. If a node's core score is inconsistent over trial blocks, then the mean core score for each node in the network is expected to be similar and thus one would expect the variance of the core scores over nodes to be small. A low variance in the mean geometrical core score over trial blocks therefore suggests either little separation between the core and periphery or an increased variability in core scores of a given node over trial blocks. Relationship Between Temporal and Geometrical Core-Periphery Organiza- tion Given the geometrical core-periphery organization in the individual layers of the multilayer networks and the temporal core-periphery organization in the full multilayer networks, it is important to ask whether brain regions in the temporal core (i.e., regions with low flexibility) are also likely to be in the geometrical core (i.e., whether they exhibit strong connectivity with other core nodes, as represented by a high value of the geometrical core score). In Fig. 6, we show scatter plots of the flexibility and core score for the 3 training levels (EXT, MOD, and MIN) and the 4 scanning sessions over the 6-week training period. We find that the temporal core-periphery organization (which is a dynamic measurement) is strongly correlated with the geometrical core score (which is a measure of network geometry and hence of 7 network structure). This indicates that regions with low temporal flexibility tend to be strongly-connected core nodes in (static) network layers. In Fig. 6, we show that the relationship between temporal and geometrical core-periphery organization occurs reliability across training depth, duration, and intensity. In Fig. 7, we show that this relationship can also be identified robustly in data extracted from individual subjects. Discussion We have shown how the mesoscale organization of functional brain networks changes over the course of learning. Our results suggest that core-periphery organization is an important and predictive component of cognitive processes that support sequential, goal-directed behavior. We summarize our findings in Fig. 7, which demonstrates that poor learners tend to have poorer separation between core and periphery (as indicated by straighter, shorter spirals in the figure) and that good learners tend to have greater separation between core and periphery (curvier, longer spirals). Our findings also demonstrate that during the generation of motor sequences, the brain consists of a temporally stable and densely connected set of core regions complemented by a temporally flexible and sparsely connected set of peripheral regions. This functional tradeoff between a core and periphery might provide a balance between the rigidity necessary to maintain motor function by the core and the adaptivity of the periphery necessary to enable behavioral change as a function of context or strategy. In the Text S1, we provide supporting results that indicate (i) that our findings are not merely a function of variation in region size and (ii) that they cannot be derived from the underlying block design of the experimental task. We also show in this supplement that (arguably) simpler properties of brain function -- such as the regional signal power of brain activity, mean connectivity strength, and parameter estimates from a general linear model -- provide less predictive power than core-periphery organization. Core-Periphery Organization of Human Brain Structure and Function. The notion of a core- periphery organization is based on the structure (rather than the temporal dynamics) of a network [30]. Intuitively, a core consists of a set of highly and mutually interconnected set of regions. In this paper, we have described what is traditionally called "core-periphery structure" using the terminology geometrical core-periphery organization. (It is geometrical rather than topological because the networks are weighted.) This intuitive notion was formalized in social networks by Borgatti and Everett in 1999 [18]. Available methods to identify and quantify geometrical core-periphery organization in networks include ones based on block models [18], k-core organization [32], and aggregation of information about connectivity and short paths through a network [33]. Unfortunately, many methods that have been used to study cores and peripheries in networks have binarized networks that are inherently weighted, which requires one to throw away a lot of important information. Even the recently developed weighted extensions of k- core decomposition [34] require a discretization of k-shells, which have been defined for both binary and weighted networks [35]. Importantly, k-core decomposition is based on a very stringent and specific type of core connectivity, so this measure misses important core-like structures [30, 36]. A well-known measure called the "rich-club coefficient" (RCC) [37] considers a different but somewhat related question of whether nodes of high degree (defined as k ≥ kt some threshold value kt) tend to connect to other nodes of high degree. (The RCC is therefore a form of assortativity.) The RCC has also been extended to weighted networks [38], but it still requires one to specify a threshold value of richness to enable one to ask whether "rich" nodes tend to connect to other "rich" nodes. The aforementioned limitations notwithstanding, several of the measures discussed above have recently been used successfully to identify a structural core of the human brain white-matter tract network, which is characterized not only by a k-core with a high value of the degree k (in particular, k ≥ 20) [34] and rich club [39,40] but also by a knotty center of nodes that have a high geodesic betweenness centrality but not 8 necessarily a high degree [36]. A k-core decomposition has also been applied to functional brain imaging data to demonstrate a relationship between network reconfiguration and errors in task performance [41]. A novel approach that is able to overcome many of these conceptual limitations is the geometrical core-score [30], which is an inherently continuous measure, is defined for weighted networks, and can be used to identify regions of a network core without relying solely on their degree or strength (i.e., weighted degree). Moreover, by using this measure, one can produce (i) continuous results, which make it possible to measure whether a brain region is more core-like or periphery-like; (ii) a discrete classification of core versus periphery; or (iii) a finer discrete division (e.g., into 3 or more groups). In addition, this method can identify multiple geometrical cores in a network and rank nodes in terms of how strongly they participate in different possible cores. This sensitivity is particularly helpful for the examination of brain networks for which multiple cores are hypothesized to mediate multimodal integration [42]. In this paper, we have demonstrated that functional brain networks derived from task-based data acquired during goal-directed brain activity exhibit geometrical core-periphery organization. Moreover, they are specifically characterized by a straightforward core-periphery landscape that includes a relatively small core composed of roughly 10% or so of the nodes in the network. In this paper, we have introduced a method and associated definitions to identify a temporal core- periphery organization based on changes in a node's module allegiance over time. We have defined the notion of a temporal core as a set of regions that exhibit fewer changes in module allegiance over time than expected in a dynamic-network null model. Neurobiologically, the temporal core contains brain areas that show consistent task-based mesoscale functional connectivity over the course of an experiment, and it is therefore perhaps unsurprising that their anatomical locations differ from nodes in the (k ≥ 20)- core [34] and RCC [39, 40] of the human white matter tract network. Our approach is inspired by the following idea: although the brain uses the function of a small subset of regions to perform a given task (i.e., some sort of core), a set of additional regions that are associated more peripherally with the task might also be activated in a transient manner. Indeed, several recent studies have highlighted the possibility of a separation between groups of regions that are consistently versus transiently activated during task-related function [43, 44], and they have demonstrated that correlations between such regions can be altered depending on their activity [43, 45]. Given the very different definitions of the geometrical and temporal cores, it is interesting that nodes in the temporal core are also likely to be present in the geometrical core. Importantly, the notions of temporal and geometrical core are complementary, and they are both intuitive in the context of brain function. A set of regions that is coherently active to perform a task (i.e., is in the geometrical core) must remain online consistently throughout an experiment (i.e., be in the temporal core), whereas a set of regions that might be activated less coherently (i.e., is in the geometrical periphery) can be utilized by separate putative functional modules over time (i.e., it can be in the temporal periphery). This interpretation is consistent with the notion that the anatomical locations of the core and periphery are task-specific. Should brain activity during other tasks also exhibit core-periphery organization, then the core and periphery of these other task networks could consist of a different set of anatomical regions than those observed here. A comparison of dynamic community structure and associated mesoscale organizational properties across brain states elicited by other tasks is outside of the scope of the present study. However, such a study in a controlled sample with similar time-series length and experimental task structure (e.g., trial lengths, block lengths, and rest periods) would likely yield important insights. Modular Versus Core-Periphery Organization. Community structure and core-periphery orga- nization are two types of mesoscale structures, and they can both be present simultaneously in a net- work [30,36]. Moreover, both modular and core-periphery organization can in principle pertain to different characteristics of or constraints on underlying brain function. In particular, the presence of community structure supports the idea of the brain containing putative functional modules, whereas the presence of a core-periphery organization underscores the fact that different brain regions likely play inherently differ- 9 ent roles in information processing. A symbiosis between these two types of organization is highlighted by the findings that we report in this manuscript: the dynamic reconfiguration of putative functional mod- ules can be described parsimoniously by temporal core-periphery organization, demonstrating that one type of mesoscale structure can help to characterize another. Furthermore, the notion that the brain can simultaneously contain functional modules (e.g., the executive network or the default-mode network) and regions that transiently mediate interactions between modules is consistent with recent characterizations of attention and cognitive control processes [46]. Dynamic Brain Networks. It is increasingly apparent that functional connectivity in the brain changes over time and that these changes are biologically meaningful. Several recent studies have high- lighted the temporal variability [47 -- 50] and non-stationarity [51] of functional brain network organization, and both of these features are apparent over short time intervals (less than 5 minutes in fMRI; less than 100 s in EEG) [47 -- 49]. Although temporal variability in functional connectivity was seen initially as a signature of measurement noise [51], recent evidence suggests instead that it might provide an indirect measurement of changing cognitive processes. Thus, it might serve as a diagnostic biomarker of dis- ease [51, 52]. Moreover, such temporal variability appears to be modulated by exogenous inputs. For example, Barnes et al. [53] demonstrated using a continuous acquisition "rest-task-rest" design that en- dogenous brain dynamics do not return to their pre-task state until approximately 18 minutes following task completion. Similar results that consider other tasks have also been reported [54]. More generally, the dynamic nature of brain connectivity is likely linked to spontaneous cortical processing, reflecting a combination of both stable and transient communication pathways [48, 49, 55]. Network Predictions of Future Learning. In this study, we observed that properties of the temporal organization of functional brain networks (e.g., on day 1 of this experiment) can be used to predict extended motor learning (e.g., on the following 10 days of home training on a discrete sequence-production task). Our findings are consistent with two previous studies that demonstrated a predictive connection between both dynamic [17] and topological [56] network organization and subsequent learning. (Note that we use the term topological because Ref. [56] considered only unweighted networks.) Reference [17] focused on early -- rather than extended -- learning of a cued sequence-production motor task (rather than a discrete one) and found that network flexibility on the first day of experiments predicted learning on the second day and that flexibility on the second day predicted learning on the third day. Reference [56] investigated participants' success in learning words of an artificial spoken language and found that network properties from individual time windows could be used to predict such success [56]. Together with the present study, these results highlight the potential breadth of the relationship between network organization and learning. The presence of such a relationship has now been identified across multiple tasks, over multiple time scales, and using both dynamic and topological network properties. Methodological Considerations. Our study has focused on large-scale changes in dynamic com- munity structure that are correlated with learning. Finer-scale investigations that employ alternative parcellation schemes [57 -- 62] with greater spatial resolution or alternative neuroimaging techniques such as EEG or MEG [55] with greater temporal resolution might uncover additional features that would enhance understanding of functional network-based predictors of learning phenomena. Throughout this paper, we have referred to feature similarities (which we estimated using the mag- nitude squared coherence) between pairs of regional BOLD time series as functional connectivity [63]. As appreciated in prior literature [64 -- 66], the interpretation of functional connectivity must be made with caution. Coherence in the activity recorded at different brain sites does not necessitate that those sites share information with one another to enable cognitive processing, as they could instead indicate that those two sites are activated by the same third party (either another brain region or an external stimulus). In this paper, we do not distinguish between these two possible drivers of strong inter-regional coherence. Future studies could employ multiple estimates of statistical associations in the form of di- agnostics [67 -- 69] and/or models [70, 71] that might uncover other sets of interactions that could predict the observed coherence structure and hence the observed behavior. 10 Materials and Methods Experiment and Data Acquisition Ethics Statement Twenty-two right-handed participants (13 females and 9 males; the mean age was about 24) volunteered with informed consent in accordance with the Institutional Review Board/Human Subjects Committee, University of California, Santa Barbara. Experiment Setup and Procedure We excluded two participants from the investigation: one participant failed to complete the experiment, and the other had excessive head motion. Our investigation therefore includes twenty participants, who all had normal/corrected vision and no history of neurological disease or psychiatric disorders. Each of these participants completed a minimum of 30 behavioral training sessions as well as 3 fMRI test sessions and a pre-training fMRI session. Training began immediately following the initial pre-training scan session. Test sessions occurred after every 2-week period of behavioral training, during which at least 10 training sessions were required. The training was done on personal laptop computers using a training module that was installed by the experimenter (N.F.W.). Participants were given instructions for how to run the module, which they were required to do for a minimum of 10 out of 14 days in a 2-week period. Participants were scanned on the first day of the experiment (scan 1), and then a second time approximately 14 days later (scan 2), once again approximately 14 days later (scan 3), and finally 14 days after that (scan 4). Not all participants were scanned exactly every two weeks; see Table S1 for details of the number of days that elapsed between scanning sessions. We asked participants to practice a set of 10-element sequences that were presented visually using a discrete sequence-production (DSP) task by generating responses to sequentially presented stimuli (see Fig. 8) using a laptop keyboard with their right hand. Sequences were presented using a horizontal array of 5 square stimuli; the responses were mapped from left to right, such that the thumb corresponded to the leftmost stimulus and the smallest finger corresponded to the rightmost stimulus. A square highlighted in red served as the imperative to respond, and the next square in the sequence was highlighted immediately following each correct key press. If an incorrect key was pressed, the sequence was paused at the error and was restarted upon the generation of the appropriate key press. Participants had an unlimited amount of time to respond and to complete each trial. All participants trained on the same set of 6 different 10-element sequences, which were presented with 3 different levels of exposure. We organized sequences so that each stimulus location was presented twice and included neither stimulus repetition (e.g., "11" could not occur) nor regularities such as trills (e.g., "121") or runs (e.g., "123"). Each training session (see Fig. 9) included 2 extensively trained sequences ("EXT") that were each practiced for 64 trials, 2 moderately trained sequences ("MOD") that were each practiced for 10 trials, and 2 minimally trained sequences ("MIN") that were each practiced for 1 trial. (See Table S1 for details of the number of trials composed of extensively, moderately, and minimally trained sequences during home training sessions.) Each trial began with the presentation of a sequence-identity cue. The purpose of the identity cue was to inform the participant what sequence they were going to have to type. For example, the EXT sequences were preceded by either a cyan (sequence A) or magenta (sequence B) circle. Participants saw additional identity cues for the MOD sequences (red or green triangles) and for the MIN sequences (orange or white stars, each of which was outlined in black). No participant reported 11 any difficulty viewing the different identity cues. Feedback was presented after every block of 10 trials; this feedback detailed the number of error-free sequences that the participant produced and the mean time it took to complete an error-free sequence. Each fMRI test session was completed after approximately 10 home training sessions (see Table S1 for details of the number of home practice sessions between scanning sessions), and each participant participated in 3 test sessions. In addition, each participant had a pre-training scan session that was identical to the other test scan sessions immediately prior to the start of training (see Fig. 9). To familiarize participants with the task, we gave a brief introduction prior to the onset of the pre-training session. We showed the participants the mapping between the fingers and the DSP stimuli, and we explained the significance of the sequence-identity cues. To help ease the transition between each participant's training environment and that of the scanner, padding was placed under his/her knees to maximize comfort. Participants made responses using a fiber-optic response box that was designed with a similar configuration of buttons as those found on the typical laptop used during training. See the lower left of Fig. 8 for a sketch of the button box used in the experiments. For instance, the center-to-center spacing between the buttons on the top row was 20 mm (compared to 20 mm from "G" to "H" on a recent MacBook Pro), and the spacing between the top row and lower left "thumb" button was 32 mm (compared to 37 mm from "G" to the spacebar on a MacBook Pro). The response box was supported using a board whose position could be adjusted to accommodate a participant's reach and hand size. Additional padding was placed under the right forearm to minimize muscle strain when a participant performed the task. Head motion was minimized by inserting padded wedges between the participant and the head coil of the MRI scanner. The number of sequence trials performed during each scanning session was the same for all participants, except for two abbreviated sessions that resulted from technical problems. In each case that scanning was cut short, participants completed 4 out of the 5 scan runs for a given session. We included data from these abbreviated sessions in this study. Participants were tested inside of the scanner with the same DSP task and the same 6 sequences that they performed during training. Participants were given an unlimited time to complete trials, though they were instructed to respond quickly but also to maintain accuracy. Trial completion was signified by the visual presentation of a fixation mark "+", which remained on the screen until the onset of the next sequence-identity cue. To acquire a sufficient number of events for each exposure type, all sequences were presented with the same frequency. Identical to training, trials were organized into blocks of 10 followed by performance feedback. Each block contained trials belonging to a single exposure type and included 5 trials for each sequence. Trials were separated by an inter-trial interval (ITI) that lasted between 0 and 6 seconds (not including any time remaining from the previous trial). Scan epochs contained 60 trials (i.e., 6 blocks) and consisted of 20 trials for each exposure type. Each test session contained 5 scan epochs, yielding a total of 300 trials and a variable number of brain scans depending on how quickly the task was performed. See Table S2 for details of the number of scans in each experimental block. Behavioral Apparatus Stimulus presentation was controlled during training using a participant's laptop computer, which was running Octave 3.2.4 (an open-source program that is very similar to Matlab) in conjunction with PsychtoolBox Version 3. We controlled test sessions using a laptop computer running Matlab version 7.1 (Mathworks, Natick, MA). We collected key-press responses and response times using a custom fiber- optic button box and transducer connected via a serial port (button box: HHSC-1 × 4-L; transducer: fORP932; Current Designs, Philadelphia, PA). 12 Behavioral Estimates of Learning Our goal was to study the relationship between brain organization and learning. To ensure independence of these two variables, we extracted brain network structure during the 4 scanning sessions, and we extracted behavioral estimates of learning in home training sessions 1 -- 10 (approximately between days 1 and 14; see Table S1), which took place before scanning session 2. For each sequence, we defined the movement time (MT) as the difference between the time of the first button press and the time of the last button press during a single sequence. For the set of sequences of a single type (i.e., sequence 1, 2, 3, 4, 5, or 6), we estimated the learning rate by fitting an exponential function (plus a constant) to the MT data [72, 73] using a robust outlier correction in Matlab (using the function fit.m in the Curve Fitting Toolbox with option "Robust" and type "Lar"): M T = D1et/κ + D2 , (1) where t is time, κ is the exponential dropoff parameter (which we call the "learning parameter") used to describe the early (and fast) rate of improvement, and D1 and D2 are real and positive constants. The sum D1+D2 is an estimate of the starting speed of a given participant prior to training, and the parameter D2 is an estimate of the fastest speed attainable by that participant after extended training. A negative value of κ indicates a decrease in MT, which is thought to indicate that learning is occurring [74, 75]. This decrease in MT has been used to quantify learning for several decades [76, 77]. Several functional forms have been suggested for the fit of MT [78, 79], and the exponential (plus constant) is viewed as the most statistically robust choice [79]. Additionally, the fitting approach that we used has the advantage of estimating the rate of learning independent of initial performance or performance ceiling. Functional MRI (fMRI) Imaging Imaging Procedures We acquired fMRI signals using a 3.0 T Siemens Trio with a 12-channel phased-array head coil. For each scan epoch, we used a single-shot echo planar imaging sequence that is sensitive to BOLD contrast to acquire 37 slices per repetition time (TR of 2000 ms, 3 mm thickness, 0.5 mm gap) with an echo time (TE) of 30 ms, a flip angle of 90 degrees, a field of view (FOV) of 192 mm, and a 64 × 64 acquisition matrix. Before the collection of the first functional epoch, we acquired a high-resolution T1-weighted sagittal sequence image of the whole brain (TR of 15.0 ms, TE of 4.2 ms, flip angle of 9 degrees, 3D acquisition, FOV of 256 mm, slice thickness of 0.89 mm, and 256 × 256 acquisition matrix). fMRI Data Preprocessing We processed and analyzed functional imaging data using Statistical Parametric Mapping (SPM8, Well- come Trust Center for Neuroimaging and University College London, UK). We first realigned raw func- tional data, then coregistered it to the native T1 (normalized to the MNI-152 template with a re-sliced resolution of 3× 3× 3 mm), and finally smoothed it using an isotropic Gaussian kernel of 8 mm full-width at half-maximum. To control for potential fluctuations in signal intensity across the scanning sessions, we normalized global intensity across all functional volumes. Network Construction Partitioning the Brain into Regions of Interest Brain function is characterized by spatial specificity: different portions of the cortex emit different, task- dependent activity patterns. To study regional specificity of the functional time series and putative interactions between brain areas, it is common to apply a standardized atlas to raw fMRI data [7, 80, 81]. 13 The choice of atlas or parcellation scheme is the topic of several recent studies in structural [57,60], resting- state [58], and task-based [59] network architecture. The question of the most appropriate delineation of the brain into nodes of a network is an open one and is guided by the particular scientific question at hand [5, 61]. Consistent with previous studies of task-based functional connectivity during learning [15, 17, 82], we parcellated the brain into 112 identifiable cortical and subcortical regions using the structural Harvard- Oxford (HO) atlas (see Table S3) installed with the FMRIB (Oxford Centre for Functional Magnetic Resonance Imaging of the Brain) Software Library (FSL; Version 4.1.1) [83, 84]. For each individual participant and each of the 112 regions, we determined the regional mean BOLD time series by separately averaging across all of the voxels in that region. Within each HO-atlas region, we constrained voxel selection to voxels that are located within an individual participant's gray matter. To do this, we first segmented each individual participant's T1 into white and gray matter volumes using the DARTEL toolbox supplied with SPM8. We then restricted the gray-matter voxels to those with an intensity of 0.3 or more (the maximum intensity was 1.0). Note that units are based on an arbitrary scale. We then spatially normalized the participant T1 and corresponding gray matter volume to the MNI-152 template -- using the standard SPM 12-parameter affine registration from the native images to the MNI-152 template image -- and resampled to 3 mm isotropic voxels. We then restricted the voxels for each HO region by using the program fslmaths [83,84] to include only voxels that are in the individual participant's gray-matter template. Wavelet Decomposition Brain function is also characterized by frequency specificity. Different cognitive and physiological func- tions are associated with different frequency bands, and this can be investigated using wavelets. Wavelet decompositions of fMRI time series have been applied extensively in both resting-state and task-based conditions [85, 86]. In both cases, they provide sensitivity for the detection of small signal changes in non-stationary time series with noisy backgrounds [87]. In particular, the maximum-overlap discrete wavelet transform (MODWT) has been used extensively in connectivity investigations of fMRI [88 -- 93]. Accordingly, we used MODWT to decompose each regional time series into wavelet scales corresponding to specific frequency bands [94]. We were interested in quantifying high-frequency components of an fMRI signal, correlations between which might be indicative of cooperative temporal dynamics of brain activity during a task. Because our sampling frequency was 2 seconds (1 TR = 2 sec), wavelet scale one provides information on the frequency band 0.125 -- 0.25 Hz and wavelet scale two provides information on the frequency band 0.06 -- 0.125 Hz. Previous work has indicated that functional associations between low-frequency components of the fMRI signal (0 -- 0.15 Hz) can be attributed to task-related functional connectivity, whereas associations between high-frequency components (0.2 -- 0.4 Hz) cannot [95]. This frequency specificity of task-relevant functional connectivity is likely due at least in part to the hemodynamic response function, which might act as a noninvertible band-pass filter on underlying neural activity [95]. Consistent with our previous work [17], we examined wavelet scale two, which is thought to be particularly sensitive to dynamic changes in task-related functional brain architecture. Construction of Dynamic Networks For each of the 112 brain regions, we extracted the wavelet coefficients of the mean time series in temporal windows given by trial blocks (of approximately 60 TRs; see Table S2). The leftmost temporal boundary of each window was equal to the first TR of an experimental trial block, and the rightmost boundary was equal to the last TR in the same block. We thereby extracted block-specific data sets from the EXT, MOD, and MIN sequences (with 6 -- 10 blocks of each sequence type; see Table S2 for details of the number 14 of blocks of each sequence type) for each of the 20 participants participating in the experiment and for each of the 4 scanning sessions. For each block-specific data set, we constructed an N × N adjacency matrix W representing the complete set of pairwise functional connections present in the brain during that window in a given participant and for a given scan. Note that N = 112 is the number of brain regions in the full brain atlas (see the earlier section on "Partitioning the Brain into Regions of Interest" for further details). To quantify the weight Wij of functional connectivity between regions labeled i and j, we used the magnitude squared spectral coherence as a measure of nonlinear functional association between any two wavelet coefficient time series (consistent with our previous study [17]). In using the coherence, which has been demonstrated to be useful in the context of fMRI neuroimaging data [95], we were able to measure frequency-specific linear relationships between time series. To examine changes in functional brain network architecture during learning, we constructed multi- layer networks by considering the set of L adjacency matrices constructed from consecutive blocks of a given sequence type (EXT, MOD, or MIN) in a given participant and scanning session. We combined the matrices in each set separately to form a rank-3 adjacency tensor A per sequence type, participant, and scan. Such a tensor can be used to represent a time-dependent network [14, 17]. In the following sections, we describe a variety of diagnostics that can be used to characterize such multilayer structures. Network Examination Dynamic Community Detection Community detection [12, 13] can be used to identify putative functional modules (i.e., sets of brain regions that exhibit similar trajectories through time). One such technique is based on the optimization of the modularity quality function [96 -- 98]. This allows one to identify groups that consist of nodes that have stronger connections among themselves than they do to nodes in other groups [12]. Recently, the modularity quality function has been generalized so that one can consider time-dependent or multiplex networks using multilayer modularity [14] Q = 1 2µXijlr {(Aijl − γlPijl) δlr + δijωjlr} δ(gil, gjr) , (2) where the adjacency matrix of layer l has components Aijl, the element Pijl gives the components of the corresponding matrix for a null model, γl is the structural resolution parameter of layer l, the quantity gil gives the community (i.e., "module") assignment of node i in layer l, the quantity gjr gives the community assignment of node j in layer r, the parameter ωjlr is the connection strength -- i.e., "interlayer coupling parameter", which gives an element of a tensor ω that constitutes a set of temporal resolution parameters if one is using the adjacency tensor A to represent a time-dependent network -- between node j in layer r and node j in layer l, the total edge weight in the network is µ = 1 layer l is κjl = kjl + cjl, the intra-layer strength of node j in layer l is kjl, and the inter-layer strength of 2Pjr κjr, the strength of node j in node j in layer l is cjl =Pr ωjlr. We employ the Newman-Girvan null model within each layer by using 2Pij Aijl is the total edge weight in layer l. We let ωjlr ≡ ω = constant for neighboring where ml = 1 layers (i.e., when l − r = 1) and ωjlr = 0 otherwise. We also let γl = γ = constant. In the main text, we report results for ω = 1 and γ = 1, and we evaluate the dependence of our results on γ and ω in the Text S1. (3) Pijl = kilkjl 2ml , Optimization of multilayer modularity (2) yields a partition of the brain regions into communities for each time window. To measure changes in the composition of communities across time (i.e., across 15 experimental blocks), we defined the flexibility fi of a node i to be the number of times that a node changed community assignment throughout the set of time windows represented by the multilayer network [17] normalized by the total number of changes that were possible (i.e., by the number of contiguous pairs of layers in the multilayer framework, which in this study ranged from 4 to 10; see Table S2). We then defined the flexibility of the entire network as the mean flexibility over all nodes in the network: F = 1 i=1 fi. To examine the relationship between brain network flexibility and learning, we confined ourselves to the two EXT (i.e., extensively trained) sequences, in which learning occurs more rapidly than in MOD and MIN sequences. We therefore estimated flexibility from the multilayer networks constructed from blocks of the two EXT sequences in the first scanning session. NPN Identification of Temporal Core, Bulk, and Periphery We find that different brain regions have different flexibilities. To determine whether a particular brain region is more or less flexible than expected, we constructed a nodal null model, which can be used to probe the individual roles of nodes in a network [15, 17]. (Note that alternative null models can be used to probe other aspects of the temporal or geometrical structure in a multilayer network [15, 17].) We rewired the ends of the multilayer network's inter-layer edges (which connect nodes in one layer to nodes in another) uniformly at random. After applying the associated permutation, an inter-layer edge can, for example, connect node i in layer t with node j 6= i in layer t + 1 rather than being constrained to connect each node i in layer t with itself in layer t + 1. We considered 100 different rewirings to construct an ensemble of 100 nodal null-model multilayer networks for each single multilayer network constructed from the brain data. We then estimated the flexibility of each node in each nodal null-model network. We created a distribution of expected mean nodal flexibility values by averaging flexibility over 100 rewirings and the 20 participants. We similarly estimated the mean nodal flexibility of the brain data by averaging flexibility over the 20 participants and 100 optimizations. (We optimized multilayer modularity using a Louvain-like locally greedy method [99, 100]. This procedure is not deterministic, so different runs of the optimization procedure can yield slightly different partitions of a network.) We considered a region to be a part of the temporal "core" if its mean nodal flexibility was below the 2.5% confidence bound of the null-model distribution, and we considered a region to be a part of the temporal "periphery" if its mean nodal flexibility was above the 97.5% confidence bound of the null-model distribution. Finally, we considered a region to be a part of the temporal "bulk" if its mean nodal flexibility was between the 2.5% and 97.5% confidence bounds of the null-model distribution. Geometrical Core-Periphery Organization To estimate the geometrical core-periphery organization of the (static) networks defined by each ex- perimental block (i.e., for each layer of a multilayer network), we used the method that was recently proposed in Ref. [30]. This method results in a "core score" (which constitutes a centrality measure) for each node that indicates where it lies on a continuous spectrum of roles between core and periphery. This method has numerous advantages over previous formulations used to study core-periphery organization. In particular, it can identify multiple geometrical cores in a network, which makes it possible to take multiple cores into account and in turn enables one to construct a detailed description of geometrical core-periphery organization by ranking the nodes in terms of how strongly they participate in different possible cores. Importantly, the continuous nature of the measure removes the need to use an artificial dichotomy of being strictly a core node versus strictly a peripheral node. In applying method, we consider a vector C with non-negative values, and we let Cij = Ci × Cj, where i and j are two nodes in an N -node network. We then seek a core vector C that satisfies the normalization condition CiCj = 1 X i,j and is a permutation of the vector C∗ whose components specify the local (geometrical) core values 16 (4) (5) C∗m = 1 1 + exp{−(m − N β) × tan(πα/2)} , m ∈ {1, . . . , N} . We seek a permutation that maximizes the core quality R =Xi,j AijCiCj . This method to compute core-periphery organization has two parameters: α ∈ [0, 1] and β ∈ [0, 1]. The parameter α sets the sharpness of the boundary between the geometrical core and the geometrical periphery. The value α = 0 yields the fuzziest boundary, and α = 1 gives the sharpest transition (i.e., a binary transition): as α varies from 0 to 1, the maximum slope of C∗ varies from 0 to +∞. The parameter β sets the size of the geometrical core: as β varies from 0 to 1, the number of nodes included in the core varies from N to 0. One now has the choice of either taking into account the local core scores of a node for a set of (α, β) coordinates sampled from [0, 1] × [0, 1] (where one weighs each choice by its corresponding value of R) or one can take into account only the score for particular choices of (α, β). Statistics and Software We performed all data analysis and statistical tests in Matlab. We performed the dynamic community detection procedure using freely available Matlab code [99] that optimizes multilayer modularity using a Louvain-like locally greedy algorithm [100]. Acknowledgements We thank Jean M. Carlson, Matthew Cieslak, John Doyle, Daniel Greenfield, and Megan T. Valentine for helpful discussions; John Bushnell for technical support; and James Fowler, Sang Hoon Lee, and Melissa Lever for comments on earlier versions of the manuscript. References 1. Saalmann YB, Pinsk MA, Wang L, Li X, Kastner S (2012) The pulvinar regulates information transmission between cortical areas based on attention demands. Science 337: 753 -- 756. 2. Josi´c K, Rubin J, Matias M, Romo R, editors (2009) Coherent Behavior in Neuronal Networks. Springer-Verlag. 3. Op de Beeck HP, Haushofer J, Kanwisher NG (2008) Interpreting fMRI data: maps, modules and dimensions. Nat Rev Neurosci 9: 123 -- 135. 4. Plaut DC (1995) Double dissociation without modularity: evidence from connectionist neuropsy- chology. J Clin Exp Neuropsych 17: 291 -- 321. 5. Bullmore ET, Bassett DS (2011) Brain graphs: graphical models of the human brain connectome. Ann Rev Clin Psych 7: 113 -- 140. 17 6. Langer N, Pedroni A, Gianotti LR, Hanggi J, Knoch D, et al. (2012) Functional brain network efficiency predicts intelligence. Hum Brain Mapp 33: 1393 -- 1406. 7. Bassett DS, Bullmore ET (2009) Human brain networks in health and disease. Curr Opin Neurol 22: 340 -- 347. 8. Matsuzaka Y, Picard N, Strick PL (2007) Skill representation in the primary motor cortex after long-term practice. J Neurophysiol 97: 1819 -- 1832. 9. Wymbs NF, Grafton ST (2009) Neural substrates of practice structure that support future off-line learning. J Neurophysiol 102: 2462 -- 2476. 10. Sun FT, Miller LM, Rao AA, D'Esposito M (2007) Functional connectivity of cortical networks involved in bimanual motor sequence learning. Cereb Cortex 17: 1227 -- 1234. 11. Mushiake H, Saito N, Sakamoto K, Itoyama Y, Tanji J (2006) Activity in the lateral prefrontal cortex reflects multiple steps of future events in action plans. Neuron 50: 631 -- 641. 12. Porter MA, Onnela JP, Mucha PJ (2009) Communities in networks. Not Amer Math Soc 56: 1082 -- 1097, 1164 -- 1166. 13. Fortunato S (2010) Community detection in graphs. Phys Rep 486: 75 -- 174. 14. Mucha PJ, Richardson T, Macon K, Porter MA, Onnela JP (2010) Community structure in time-dependent, multiscale, and multiplex networks. Science 328: 876 -- 878. 15. Bassett DS, Porter MA, Wymbs NF, Grafton ST, Carlson JM, et al. (2012) Robust detection of dynamic community structure in networks. Chaos 23: 013142. 16. Holme P, Saramaki J (2012) Temporal networks. Phys Rep 519: 97 -- 125. 17. Bassett DS, Wymbs NF, Porter MA, Mucha PJ, Carlson JM, et al. (2011) Dynamic reconfiguration of human brain networks during learning. Proc Natl Acad Sci USA 108: 7641 -- 7646. 18. Borgatti SP, Everett MG (1999) Models of core/periphery structures. Social Networks 21: 375 -- 395. 19. Dodge Y (2003) The Oxford Dictionary of Statistical Terms. Oxford University Press. 20. Darlington RB (1970) Is kurtosis really 'peakedness' ? The American Statistician 24: 19 -- 22. 21. Mesulam MM (1998) From sensation to cognition. Brain 121: 1013 -- 1052. 22. Bischoff-Grethe A, Goedert KM, Willingham DT, Grafton ST (2004) Neural substrates of response-based sequence learning using fMRI. J Cogn Neurosci 16: 127 -- 138. 23. Picard N, Strick PL (1997) Activation on the medial wall during remembered sequences of reaching movements in monkeys. J Neurophysiol 77: 2197 -- 2201. 24. Kennerley SW, Sakai K, Rushworth MF (2004) Organization of action sequences and the role of the pre-SMA. J Neurophysiol 91: 978 -- 993. 25. Willingham DB, Salidis J, Gabrieli JD (2002) Direct comparison of neural systems mediating conscious and unconscious skill learning. J Neurophysiol 88: 1451 -- 1460. 18 26. Destrebecqz A, Peigneux P, Laureys S, Degueldre C, Del Fiore G, et al. (2005) The neural cor- relates of implicit and explicit sequence learning: Interacting networks revealed by the process dissociation procedure. Learn Mem 12: 480 -- 490. 27. Brown R, Robertson E (2007) Inducing motor skill improvements with a declarative task. Nat Neurosci 10: 148 -- 149. 28. Guimer`a R, Amaral LAN (2005) Functional cartography of complex metabolic networks. Nature 433: 895 -- 900. 29. Guimer`a R, Sales-Pardo M, Amaral LAN (2007) Classes of complex networks defined by role-to- role connectivity profiles. Nat Phys 3: 63 -- 69. 30. Rombach MP, Porter MA, Fowler JH, Mucha PJ (2012) Core-periphery structure in networks. SIAM Journal of Applied Mathematics In Press. 31. Opsahl T, Agneessens F, Skvoretz J (2010) Node centrality in weighted networks: Generalizing degree and shortest paths. Social Networks 32: 245 -- 251. 32. Holme P (2005) Core-periphery organization of complex networks. Phys Rev E 72: 046111. 33. da Silva MR, Ma H, Zeng AP (2008) Centrality, network capacity, and modularity as parameters to analyze the core-periphery structure in metabolic networks. Proceedings of the IEEE 96: 1411-1420. 34. Hagmann P, Cammoun L, Gigandet X, Meuli R, Honey CJ, et al. (2008) Mapping the structural core of human cerebral cortex. PLoS Biol 6: e159. 35. Garas A, Schweitzer F, Havlin S (2012) A k-shell decomposition method for weighted networks. New J Phys 14. 36. Shanahan M, Wildie M (2012) Knotty-centrality: finding the connective core of a complex net- work. PLoS One 7: e36579. 37. Colizza V, Flammini A, Serrano MA, Vespignani A (2006) Detecting rich-club ordering in complex networks. Nat Phys 2: 110 -- 116. 38. Opsahl T, Colizza V, Panzarasa P, Ramasco JJ (2008) Prominence and control: the weighted rich-club effect. Phys Rev Lett 101: 168702. 39. van den Heuvel MP, Sporns O (2011) Rich-club organization of the human connectome. J Neurosci 31: 15775 -- 15786. 40. van den Heuvel MP, Kahn RS, Goni J, Sporns O (2012) High-cost, high-capacity backbone for global brain communication. Proc Natl Acad Sci USA 109: 11372 -- 11377. 41. Ekman M, Derrfuss J, Tittgemeyer M, Fiebach CJ (2012) Predicting errors from reconfiguration patterns in human brain networks. Proc Natl Acad Sci U S A 109: 16714 -- 16719. 42. Sepulcre J, Sabuncu MR, Yeo TB, Liu H, Johnson KA (2012) Stepwise connectivity of the modal cortex reveals the multimodal organization of the human brain. J Neurosci 32: 10649-10661. 43. Fornito A, Harrison BJ, Zalesky A, Simons JS (2012) Competitive and cooperative dynamics of large-scale brain functional networks supporting recollection. Proc Natl Acad Sci USA 109: 12788 -- 12793. 19 44. Dosenbach NU, Visscher KM, Palmer ED, Miezin KK F M andd Wenger, Kang HC, et al. (2006) A core system for the implementation of task sets. Neuron 50: 799 -- 812. 45. de Pasquale F, Della Penna S, Snyder AZ, Marzetti L, Pizzella V, et al. (2012) A cortical core for dynamic integration of functional networks in the resting human brain. Neuron 74: 753 -- 764. 46. Uddin LQ, Supekar KS, Ryali S, Menon V (2011) Dynamic reconfiguration of structural and functional connectivity across core neurocognitive brain networks with development. J Neurosci 31: 18578 -- 18589. 47. Whitlow CT, Casanova R, Maldjian JA (2011) Effect of resting-state functional mr imaging duration on stability of graph theory metrics of brain network connectivity. Radiology 259: 516 -- 524. 48. Kramer MA, Eden UT, Lepage KQ, Kolaczyk ED, Bianchi MT, et al. (2011) Emergence of persistent networks in long-term intracranial eeg recordings. J Neurosci 31: 15757 -- 15767. 49. Chu CJ, Kramer MA, Pathmanathan J, Bianchi MT, Westover MB, et al. (2012) Emergence of stable functional networks in long-term human electroencephalography. J Neurosci 32: 2703 -- 2713. 50. Allen EA, Damaraju E, Plis SM, Erhardt EB, Eichele T, et al. (2012) Tracking whole-brain connectivity dynamics in the resting state. Cereb Cortex Epub ahead of print. 51. Jones DT, Vemuri P, Murphy MC, Gunter JL, Senjem ML, et al. (2012) Non-stationarity in the "resting brain's" modular architecture. PLoS One 7: e39731. 52. Siebenhuhner F, Weiss SA, Coppola R, Weinberger DR, Bassett DS (2012) Intra- and inter- frequency brain network structure in health and schizophrenia. arXiv 1209.0729. 53. Barnes A, Bullmore ET, Suckling J (2009) Endogenous human brain dynamics recover slowly following cognitive effort. PLoS One 4: e6626. 54. Tambini A, Ketz N, Davachi L (2010) Enhanced brain correlations during rest are related to memory for recent experiences. Neuron 65: 280 -- 290. 55. Doron KW, Bassett DS, Gazzaniga MS (2012) Dynamic network structure of interhemispheric coordination. Proc Natl Acad Sci U S A 109: 18661 -- 18668. 56. Sheppard JP, Wang JP, Wong PC (2012) Large-scale cortical network properties predict future sound-to-word learning success. J Cogn Neurosci 24: 1087 -- 1103. 57. Zalesky A, Fornito A, Harding IH, Cocchi L, Yucel M, et al. (2010) Whole-brain anatomical networks: Does the choice of nodes matter? NeuroImage 50: 970 -- 983. 58. Wang J, Wang L, Zang Y, Yang H, Tang H, et al. (2009) Parcellation-dependent small-world brain functional networks: A resting-state fMRI study. Hum Brain Mapp 30: 1511 -- 1523. 59. Power JD, Cohen AL, Nelson SM, Wig GS, Barnes KA, et al. (2011) Functional network organi- zation of the human brain. Neuron 72: 665 -- 678. 60. Bassett DS, Brown JA, Deshpande V, Carlson JM, Grafton ST (2011) Conserved and variable architecture of human white matter connectivity. NeuroImage 54: 1262 -- 1279. 61. Wig GS, Schlaggar BL, Petersen SE (2011) Concepts and principles in the analysis of brain networks. Ann NY Acad Sci 1224: 126 -- 146. 20 62. de Reus MA, van den Heuvel MP (2013) The parcellation-based connectome: Limitations and extensions. NeuroImage Epub ahead of print. 63. Friston KJ (2011) Functional and effective connectivity: a review. Brain Connect 1: 13 -- 36. 64. Horwitz B (2003) The elusive concept of brain connectivity. NeuroImage 19: 466 -- 470. 65. Gavrilescu M, Stuart GW, Rossell S, Henshall K, McKay C, et al. (2008) Functional connectivity estimation in fMRI data: influence of preprocessing and time course selection. Hum Brain Mapp 29: 1040 -- 1052. 66. Jones TB, Bandettini PA, Kenworthy L, Case LK, Milleville SC, et al. (2010) Sources of group differences in functional connectivity: an investigation applied to autism spectrum disorder. Neu- roImage 49: 401 -- 414. 67. David O, Cosmelli D, Friston KJ (2004) Evaluation of different measures of functional connectivity using a neural mass model. NeuroImage 21: 659 -- 673. 68. Smith SM, Miller KL, Salimi-Khorshidi G, Webster M, Beckmann CF, et al. (2011) Network modelling methods for FMRI. NeuroImage 54: 875 -- 91. 69. Ryali S, Chen T, Supekar K, Menon V (2012) Estimation of functional connectivity in fMRI data using stability selection-based sparse partial correlation with elastic net penalty. NeuroImage 59: 3852 -- 3861. 70. Gates KM, Molenaar PC (2012) Group search algorithm recovers effective connectivity maps for individuals in homogeneous and heterogeneous samples. NeuroImage 63: 310 -- 319. 71. Watanabe T, Hirose S, Wada H, Imai Y, Machida T, et al. (2013) A pairwise maximum entropy model accurately describes resting-state human brain networks. Nat Comm 4: 1370. 72. Schmidt RA, Lee TD (2005) Motor Control and Learning: A Behavioral Emphasis. Human Kinetics, fourth edition. 73. Rosenbaum DA (2010) Human Motor Control. Elsevier. 74. Yarrow K, Brown P, Krakauer JW (2009) Inside the brain of an elite athlete: the neural processes that support high achievement in sports. Nat Rev Neurosci 10: 585 -- 596. 75. Dayan E, Cohen LG (2011) Neuroplasticity subserving motor skill learning. Neuron 72: 443 -- 454. 76. Snoddy GS (1926) Learning and stability: A psychophysical analysis of a case of motor learning with clinical applications. J App Psych 10: 1 -- 36. 77. Crossman ERFW (1959) A theory of the acquisition of speed-skill. Ergonomics 2: 153 -- 166. 78. Newell KM, Rosenbloom PS (1981) Mechanisms of skill acquisition and the law of practice. In: Anderson JR, editor, Cognitive skills and their acquisition, Lawrence Erlbaum Associates. pp. 1 -- 55. 79. Heathcote A, Brown S, Mewhort DJ (2000) The power law repealed: the case for an exponential law of practice. Psychon Bull Rev 7: 185 -- 207. 80. Bassett DS, Bullmore ET (2006) Small-world brain networks. Neuroscientist 12: 512 -- 523. 81. Bullmore E, Sporns O (2009) Complex brain networks: Graph theoretical analysis of structural and functional systems. Nat Rev Neurosci 10: 186 -- 198. 21 82. Mantzaris AV, Bassett DS, Wymbs NF, Estrada E, Porter MA, et al. (2013) Dynamic network centrality summarizes learning in the human brain. Journal of Complex Networks Epub Ahead of Print. 83. Smith SM, Jenkinson M, Woolrich MW, Beckmann CF, Behrens TEJ, et al. (2004) Advances in functional and structural MR image analysis and implementation as FSL. NeuroImage 23: 208 -- 219. 84. Woolrich MW, Jbabdi W, Patenaude B, Chappell M, Makni S, et al. (2009) Bayesian analysis of neuroimaging data in FSL. NeuroImage 45: S173 -- S186. 85. Bullmore E, Fadili J, Breakspear M, Salvador R, Suckling J, et al. (2003) Wavelets and statistical analysis of functional magnetic resonance images of the human brain. Stat Methods Med Res 12: 375 -- 399. 86. Bullmore E, Fadili J, Maxim V, Sendur L, Whitcher B, et al. (2004) Wavelets and functional magnetic resonance imaging of the human brain. NeuroImage 23: S234 -- S249. 87. Brammer MJ (1998) Multidimensional wavelet analysis of functional magnetic resonance images. Hum Brain Mapp 6: 378 -- 382. 88. Achard S, Salvador R, Whitcher B, Suckling J, Bullmore E (2006) A resilient, low-frequency, small-world human brain functional network with highly connected association cortical hubs. J Neurosci 26: 63 -- 72. 89. Bassett DS, Meyer-Lindenberg A, Achard S, Duke T, Bullmore E (2006) Adaptive reconfiguration of fractal small-world human brain functional networks. Proc Natl Acad Sci USA 103: 19518 -- 19523. 90. Achard S, Bullmore E (2007) Efficiency and cost of economical brain functional networks. PLoS Comp Biol 3: e17. 91. Achard S, Bassett DS, Meyer-Lindenberg A, Bullmore E (2008) Fractal connectivity of long- memory networks. Phys Rev E 77: 036104. 92. Bassett DS, Meyer-Lindenberg A, Weinberger DR, Coppola R, Bullmore E (2009) Cognitive fitness of cost-efficient brain functional networks. Proc Natl Acad Sci USA 106: 11747 -- 11752. 93. Lynall ME, Bassett DS, Kerwin R, McKenna P, Muller U, et al. (2010) Functional connectivity and brain networks in schizophrenia. J Neurosci 30: 9477 -- 9487. 94. Percival DB, Walden AT (2000) Wavelet Methods for Time Series Analysis. Cambridge University Press. 95. Sun FT, Miller LM, D'Esposito M (2004) Measuring interregional functional connectivity using coherence and partial coherence analyses of fMRI data. NeuroImage 21: 647 -- 658. 96. Newman MEJ, Girvan M (2004) Finding and evaluating community structure in networks. Phys Rev E 69: 026113. 97. Newman MEJ (2006) Modularity and community structure in networks. Proc Natl Acad Sci USA 103: 8577 -- 8582. 98. Newman MEJ (2006) Finding community structure in networks using the eigenvectors of matrices. Phys Rev E 74: 036104. 22 99. Jutla IS, Jeub LGS, Mucha PJ (2011 -- 2012). A generalized Louvain method for community detection implemented in matlab. URL http://netwiki.amath.unc.edu/GenLouvain. 100. Blondel VD, Guillaume JL, Lambiotte R, Lefebvre E (2008) Fast unfolding of community hierar- chies in large networks. J Stat Mech : P10008. 101. Matsumoto M, Nishimura T (1998) Mersenne twister: a 623-dimensionally equidistributed uni- form pseudo-random number generator. ACM Transactions on Modeling and Computer Simula- tion 8: 3 -- 30. Figure and Table Legends Figure 1. Network Organization of Human Brain Dynamics. (A) Temporal Networks of the Human Brain. We parcellate the brain into anatomical regions that can be represented as nodes in a network, and we use the coherence between functional Magnetic Resonance Imaging (fMRI) time series of each pair of nodes over a time window to determine the weight of the network edge connecting those nodes. We determine these weights separately using approximately 10 non-overlapping time windows of 2 -- 3 min duration and thereby construct temporal networks that represent the dynamical functional connectivity in the brain. (B) Cohesive Mesoscale Structures. (top) An example of a network with a modular organization in which high-degree nodes (brown) are often found in the center of modules or bridging distinct modules that are composed mostly of low-degree nodes (blue). (bottom) A network with a core-periphery organization in which nodes in the core (purple) are more densely connected with one another than nodes in the periphery are with one another (green). A brain network nodes 20 0 −20 40 0 20 0 −20 −40 0 nodal time series B modular organization 20 40 60 80 100 20 40 60 80 100 time windows . r=0.20 r=0.29 r=0.25 r=0.60 . . . core-periphery organization pairwise coherence temporal networks 23 Figure 2. Dynamic Network Diagnostics Change with Learning. (A) Multilayer modularity, (B) number of communities, and (C) mean flexibility calculated as a function of the number of trials completed after a scanning session (see Table 1 for the relationship between the number of trials practiced and training duration and intensity). We average the values for each diagnostic over the 100 multilayer modularity optimizations, and we average flexibility over the 112 brain regions (in addition to averaging over the 100 optimizations per subject). Error bars indicate the standard error of the mean over participants. Figure 3. Temporal Core-Periphery Organization of the Brain determined using fMRI signals during the performance of a simple motor learning task. (A) The core (cyan), bulk (gold), and periphery (maroon) nodes consist, respectively, of brain regions whose mean flexibility over individuals is less than, equal to, and greater than that expected in a null model (gray shaded region). We measure flexibility based on the allegiance of nodes to putative functional modules. Error bars indicate the standard error of the mean over individuals. (B) The anatomical distribution of regions in the core, bulk, and periphery appears to be spatially contiguous. The core primarily contains sensorimotor and visual processing areas, the periphery primarily contains multimodal association areas, and the bulk contains the remainder of the brain (and is therefore composed predominantly of frontal and temporal cortex). 102 number of trials 103 C 0.15 0.14 0.13 0.12 0.11 0.1 0.09 (cid:31)exibility 102 number of trials 103 4 3.9 3.8 3.7 3.6 3.5 3.4 3.3 3.2 number of communities B 102 number of trials 103 0.094 0.092 0.09 0.088 0.086 0.084 0.082 0.08 0.078 multilayer modularity A periphery bulk core B null model 0.1 flexibility 0.15 A 100 80 60 40 20 brain region 0 0.05 24 Figure 4. Geometrical Core-Periphery Organization in Brain Networks. (A) Core quality R (5) in the (α, β) parameter plane for a typical participant (3), scanning session (1), sequence type (EXT), and experimental block (1). (B) Distribution of the α and β values that maximize the R-score. We compute this distribution over all network layers, participants, scanning sessions, and sequence types. The β parameter is much more localized (its standard deviation is 0.05) than the α parameter (its standard deviation is 0.26). (C) Mean core shape. We plot the ordered vector of C values. We have set the values of α and β to the mean values of those that maximize the R-score for all network layers, participants, scanning sessions, and sequence types. Figure 5. Geometrical Core Scores Change with Learning. Variance of the distribution of mean geometrical core scores over brain regions as a function of the number of trials completed after a scanning session. (See Table 1 for the relationship between the number of trials practiced and training duration and intensity.) Error bars indicate the standard error of the mean over participants (where the data point from each participant is the mean geometrical core score over brain regions, scanning sessions, sequence types, and network layers). 40 brain region 60 80 100 1 0.8 0.6 0.4 0.2 core scoreC 0.2 0.4 α 0.6 0.8 1 0.2 0.4 β 0.6 0.8 1 0 0 20 0.2 0.1 0 0 0.6 0.4 0.2 0 0 relative frequency B core quality 0.6 0.5 0.4 0.3 0.2 .1 00 0.2 0.4 0.6 0.8 β 1 A α 0.2 0.4 0.6 0.8 1 102 number of trials 103 0.015 0.01 0.005 variance of geometrical core score 25 .= Figure 6. Relationship Between Temporal and Geometrical Core-Periphery Organizations. A strong negative correlation exists between flexibility and the geometrical core score for networks constructed from blocks of (A) extensively, (B) moderately, and (C) minimally trained sequences on scanning session 1 (day 1; circles), session 2 (after approximately 2 weeks of training; squares), session 3 (after approximately 4 weeks of training; diamonds), and session 4 (after approximately 6 weeks of training; stars). This negative correlation indicates that the temporal core-periphery organization is mimicked in the geometrical core-periphery organization and therefore that the core of dynamically stiff regions also exhibits dense connectivity. We show temporal core nodes in cyan, temporal bulk nodes in gold, and temporal periphery nodes in maroon. The darkness of data points indicates scanning session; darker colors indicate earlier scans, so the darkest colors indicate scan 1 and the lightest ones indicate scan 4. The grayscale lines indicate the best linear fits; again, darker colors indicate earlier scans, so session 1 is in gray and session 4 is in light gray. The Pearson correlation between the flexibility (averaged over 100 multilayer modularity optimizations, 20 participants, and 4 scanning sessions) and the geometrical core score (averaged over 20 participants and 4 scanning sessions) is significant for the EXT (r p 3.4 × 10−45), MOD (r −0.93, .= .= .= −0.92, p 4.8 × 10−50) data. .= .= 2.2 × 10−49), and MIN (r −0.93, p minimal training 0.4 0.3 0.2 0.1 0 −0.1 0.05 0.1 flexibility 0.15 C geometrical core score moderate training 0.05 0.1 flexibility 0.15 0.4 0.3 0.2 0.1 0 −0.1 geometrical core score B scan1 scan 2 scan 3 scan 4 Core Bulk Periphery 0.4 0.3 0.2 0.1 0 geometrical core score −0.1 0.05 0.1 flexibility 0.15 A extensive training 26 Figure 7. Core-Periphery Organization of Brain Dynamics During Learning. The relationship between temporal and geometrical core-periphery organization and their associations with learning are present in individual subjects. We represent this relationship using spirals in a plane; data points in this plane represent brain regions located at the polar coordinates (f s,−f κ), where f is the flexibility of the region, s is the skewness of flexibility over all regions, and κ is the learning parameter (see the Materials and Methods) that describes each individual's relative improvement between sessions. The skewness predicts individual differences in learning; the Spearman rank correlation is ρ and p good learners (curvier spirals) tend to have high skewness (long spirals). Color indicates flexibility: blue nodes have lower flexibility, and brown nodes have higher flexibility. −0.480 0.034. Poor learners (straighter spirals) tend to have a low skewness (short spirals), whereas .= .= f s -fκ 0 1 2 −1 −2 −3 −4 −2 −1 0 1 2 3 4 0.14 0.13 0.12 0.11 (cid:30)exibility 0.1 0.09 0.08 0.07 0.06 0.05 geometrical core score 27 Figure 8. Trial Structure and Stimulus-Response (S-R) Mapping. (A) Each trial began with the presentation of a sequence-identity cue that remained on screen for 2 seconds. Each of the 6 trained sequences was paired with a unique identity cue. A discrete sequence-production (DSP) event structure was used to guide sequence production. The onset of the initial DSP stimulus (thick square, colored red in the task) served as the imperative to produce the sequence. A correct key press led to the immediate presentation of the next DSP stimulus (and so on) until the 10-element sequence was correctly executed. Participants received a feedback "+" to signal that a sequence was completed and to wait (approximately 0 -- 6 seconds) for the start of the next trial. This waiting period is called the "inter-trial interval" (ITI). At any point, if an incorrect key was hit, a participant would receive an error signal (not shown in the figure) and the DSP sequence would pause until the correct response was received. (B) There was a direct S-R mapping between a conventional keyboard or an MRI-compatible button box (see the lower left of the figure) and a participant's right hand, so the leftmost DSP stimulus cued the thumb and the rightmost stimulus cued the pinky finger. Note that the button location for the thumb was positioned to the lower left to achieve maximum comfort and ease of motion. a) A sequence identity cue B 1 1 thmb 2 0 0 DSP task production 0 m s 2 S-R mapping indx mdl rng pnky 2 3 4 5 3 2 4 5 Feedback + 0 0 0 m s + v a r i a b l e I T I 28 Figure 9. Experiment Timeline. Training sessions in the MRI scanner during the collection of blood-oxygen-level-dependent (BOLD) signals were interleaved with training sessions at home. Participants first practiced the sequences in the MRI scanner during a baseline training session (top). Following every approximately 10 training sessions (see Table S1), participants returned for another scanning session. During each scanning session, a participant practiced each sequence for 50 trials. Participants trained at home between the scanning sessions (bottom). During each home training session, participants practiced the sequences in a random order. (We determined a random order using the Mersenne Twister algorithm of Nishimura and Matsumoto [101] as implemented in the random number generator rand.m of Matlab version 7.1). Each EXT sequence was practiced for 64 trials, each MOD sequence was practiced for 10 trials, and each MIN sequence was practiced for 1 trial. Figure 10. (Supplementary Figure S1) Reliability of Temporal Core-Periphery Structure. Temporal core (cyan), bulk (gold), and periphery (maroon) of dynamic networks determined based on the flexibility of trial blocks in which participants practiced sequences that would eventually be extensively trained. (A) Flexibility of the temporal core, bulk, and periphery averaged over the 100 multilayer modularity optimizations and 20 participants for blocks composed of extensively trained (EXT; light circles), moderately trained (MOD; squares), and minimally trained (MIN; dark diamonds) sequences. The darkness of data points indicates scanning session; darker colors indicate earlier scans, so the darkest colors indicate scan 1 and the lightest ones indicate scan 4. (B) The coefficient of variation of flexibility calculated over the 100 optimizations and 3 sequence types for all brain regions. Error bars indicate the standard error of the mean CV over participants. Both panels use data from scanning session 1 on day 1 of the experiment (which is prior to home training). experiment timeline training sessions 1-10 training sessions 11-20 training sessions 21-30 SCAN SCAN SCAN SCAN "baseline training" (50 trials/sequence) (50 trials/sequence) (50 trials/sequence) home training example EXT (64 trials /sequence) MOD (10 trials /sequence) MIN (1 trial /sequence) 29 Figure 11. (Supplementary Figure S2) Temporal Core-Periphery Organization Over 42 Days. Temporal core (cyan), bulk (gold), and periphery (maroon) of dynamic networks defined by trial blocks in which participants practiced sequences that would eventually be (A) extensively trained, (B) moderately trained, and (C) minimally trained for data from scanning sessions 2 (after approximately 2 weeks of training; circles), 3 (after approximately 4 weeks of training; squares), and 4 (after approximately 6 weeks of training; diamonds). The darkness of data points indicates scanning session; darker colors indicate earlier scans, so the darkest colors indicate scan 1 and the lightest ones indicate scan 4. Figure 12. (Supplementary Figure S3) Geometrical Core-Periphery Organization Over 42 Days. Geometrical core scores for each brain region defined by the trial blocks in which participants practiced sequences that would eventually be (A) extensively trained, (B) moderately trained, and (C) minimally trained for data from scanning sessions 1 (day 1; black circles), 2 (after approximately 2 weeks of training; dark gray squares), 3 (after approximately 4 weeks of training; gray diamonds), and 4 (after approximately 6 weeks of training; light gray stars). We have averaged the geometrical core scores over blocks and over 20 participants. The order of brain regions is identical for all 3 panels (A-C ), and we chose this order by ranking regions from high to low geometrical core scores from the EXT blocks on scanning session 1 (on day 1 of the experiment). Figure 13. (Supplementary Figure S4) Relationship Between Temporal Core-Periphery Organization and Community Structure. (A) Mean-coherence matrix over all EXT blocks from all participants on scanning day 1. The colored bars above the matrix indicate the 3 communities that we identified from the representative partition. Mean partition similarity z-score zi over all participants for blocks of (B) extensively, (C) moderately, and (D) minimally trained sequences for all 4 scanning sessions over the approximately 6 weeks of training. The horizontal gray lines in panels (B-D) indicate the zi value that corresponds to a right-tailed p-value of 0.05. 0.92. (B) Box plot over .= Figure 14. (Supplementary Figure S5) Region Size is Uncorrelated with Flexibility. (A) Scatter plot of the size of the brain region in voxels (averaged over participants) versus the flexibility of the EXT multilayer networks, which we averaged over the 100 multilayer modularity optimizations and the 20 participants. Data points indicate brain regions. The line indicates the best linear fit. Its Pearson correlation coefficient is r the 20 participants of the squared Pearson correlation coefficient r2 between the participant-specific region size in voxels and the participant-specific flexibility averaged over the 100 multilayer modularity optimizations. −0.009, and the associated p-value is p .= 30 Figure 15. (Supplementary Figure S6) Temporal Core-Periphery Organization and Task-Related Activations. Mean GLM parameter estimates for the temporal core (cyan; circles), bulk (gold; squares), and periphery (maroon; diamonds) of dynamic networks defined by the trial blocks in which participants practiced sequences that would eventually be (A) extensively trained, (B) moderately trained, and (C) minimally trained for data from scanning sessions 1 (first day of training), 2 (after approximately 2 weeks of training), 3 (after approximately 4 weeks of training), and 4 (after approximately 6 weeks of training). Figure 16. (Supplementary Figure S7) Effect of Structural Resolution Parameter. (A,B) Number of communities and (C,D) number of regions in the temporal core (cyan; circles), temporal bulk (gold; squares), and temporal periphery (maroon; diamonds) as a function of the structural resolution parameter γ, where we considered (A,C) γ ∈ [0.2, 5] in increments of ∆γ = 0.2 and (B,D) γ ∈ [0.8, 1.8] in increments of ∆γ = 0.01. We averaged the values in panels (A) and (B) over 100 multilayer modularity optimizations and over the 20 participants. Figure 17. (Supplementary Figure S8) Effect of Temporal Resolution Parameter. (A) Number of communities averaged over 100 multilayer modularity optimizations and over 20 participants as a function of the temporal resolution parameter ω. (B) Number of regions that we identified as part of the temporal core (cyan; circles), temporal bulk (gold; squares), and temporal periphery (maroon; diamonds) as we vary ω from 0.1 to 2 in increments of ∆ω = 0.1. 31 (Main Text) Table 1: Relationship Between Training Duration, Intensity, and Depth. We report the number of trials (i.e., "depth") of each sequence type (i.e., "intensity") completed after each scanning session (i.e., "duration") averaged over the 20 participants. (Supplementary Table S1): Experimental Details for Behavioral Data Acquired Between Scanning Sessions. We give the minimum, mean, maximum, and standard error of the mean over participants for the following variables: the number of days between scanning sessions; the number of practice sessions performed at home between scanning sessions; and the number of trials composed of extensively, moderately, and minimally trained sequences during home practice between scanning sessions. (Supplementary Table S2): Experimental Details for Brain Imaging Data Acquired Dur- ing Scanning Sessions. In the top three rows, we give the mean, minimum, maximum, and standard error over participants for the number of blocks composed of extensively, moderately, and minimally trained sequences during scanning sessions. In the bottom three rows, we give (in TRs) the mean, mini- mum, maximum, and standard error of the length over blocks composed of extensively, moderately, and minimally trained sequences during scanning sessions. (Supplementary Table S3): Brain regions in the Harvard-Oxford (HO) Cortical and Subcortical Parcellation Scheme provided by FSL [83, 84] and their affiliation to the temporal core (C; cyan), bulk (B; gold), and periphery (P; maroon) for both left (L) and right (R) hemispheres. Session 1 Session 2 Session 3 Session 4 MIN Sequences MOD Sequences EXT Sequences 50 50 50 110 200 740 170 350 1430 230 500 2120 Table 1. Relationship Between Training Duration, Intensity, and Depth. We report the number of trials (i.e., "depth") of each sequence type (i.e., "intensity") completed after each scanning session (i.e., "duration") averaged over the 20 participants. Supplementary Material for "Core-Periphery Organization of Human Brain Dynamics" Danielle S. Bassett1,2,∗, Nicholas F. Wymbs5, M. Puck Rombach3,4, Mason A. Porter3,4, Peter J. Mucha6,7, Scott T. Grafton5 1Department of Physics, University of California, Santa Barbara, CA 93106, USA; 2 Sage Center for the Study of the Mind, University of California, Santa Barbara, CA 93106; 3 Oxford Centre for Industrial and Applied Mathematics, Mathematical Institute, University of Oxford, Oxford OX1 3LB, UK; 4CABDyN Complexity Centre, University of Oxford, Oxford, OX1 1HP, UK; 5Department of Psychological and Brain Sciences and UCSB Brain Imaging Center, University of California, Santa Barbara, CA 93106, USA; 6Carolina Center for Interdisciplinary Applied Mathematics, Department of Mathematics, University of North Carolina, Chapel Hill, NC 27599, USA; 7Institute for Advanced Materials, Nanoscience & Technology, University of North Carolina, Chapel Hill, NC 27599, USA; ∗Corresponding author. Email address: [email protected] October 30, 2018 arXiv:1210.3555v2 [q-bio.NC] 30 Oct 2013 1 Figure 1: Reliability of Temporal Core-Periphery Structure. Temporal core (cyan), bulk (gold), and periphery (maroon) of dynamic networks determined based on the flexibility of trial blocks in which participants practiced sequences that would eventually be extensively trained. (A) Flexibility of the temporal core, bulk, and periphery averaged over the 100 multilayer modularity optimizations and 20 participants for blocks composed of extensively trained (EXT; light circles), moderately trained (MOD; squares), and minimally trained (MIN; dark diamonds) sequences. The darkness of data points indicates scanning session; darker colors indicate earlier scans, so the darkest colors indicate scan 1 and the lightest ones indicate scan 4. (B) The coefficient of variation of flexibility calculated over the 100 optimizations and 3 sequence types for all brain regions. Error bars indicate the standard error of the mean CV over participants. Both panels use data from scanning session 1 on day 1 of the experiment (which is prior to home training). Reliability of Temporal Core-Periphery Organization A brain region's role in the temporal core, bulk, and periphery is robust across levels of training. Regions identified as part of the core, bulk, or periphery in multilayer networks constructed from the EXT blocks in scanning session 1 have similar flexibilities in the other two levels of training (MOD and MIN; see Fig. 1A) for the same scanning session. To quantify the variability of a brain region's flexibility, we calculated the coefficient of variation (CV) of flexibility over the 100 optimizations and the 3 levels of training (see Fig. 1B). The CV is defined as CV = σ/µ, where σ is the standard deviation of a given sample and µ is its mean. We observe that the variabilities over optimizations and scans (i.e., CV) and over participants (i.e., error bars) are largest in regions designated as part of the temporal core and smallest in regions designated as part of the temporal periphery. In addition, regional flexibility is also conserved across both intensity of training (MIN, MOD, and EXT) and duration of training (sessions 1 -- 4). Observe in Fig. 2 that regions identified as part of the temporal core in multilayer networks constructed from the EXT blocks in scanning session 1 exhibit small flexibility for all other scanning sessions and for all 3 training levels (EXT, MOD, and MIN). Regions in the temporal bulk and temporal periphery exhibit a similar amount of flexibility to one another. Reliability of Geometrical Core-Periphery Organization As we illustrate in Fig. 3, the geometrical core-periphery organization of the brain was consistent over the 42 days of practice, across sequence types, and throughout variations in the intensity of training (MIN, MOD, and EXT) and in the duration of training (sessions 1 -- 4). Relationship Between Temporal Core-Periphery Organization and Community Structure The division of the brain networks into temporal core, bulk, and peripheral nodes has interesting similarities to their partitioning into communities based on optimizing multilayer modularity. We first noted this sim- ilarity when we examined community structure in an object that we call the mean-coherence matrix. The mean-coherence matrix ¯A contain elements ¯Aij that are equal to the mean coherence between nodes i and 2 scan 1 core bulk periphery 40 80 brain region 120 2.5 2 1.5 1 0.5 0 0 coefficient of variation scan 1 B core bulk periphery 40 80 brain region 120 0.16 0.12 0.08 0.04 0 A flexibility extensive moderate minimal Figure 2: Temporal Core-Periphery Organization Over 42 Days. Temporal core (cyan), bulk (gold), and periphery (maroon) of dynamic networks defined by trial blocks in which participants practiced sequences that would eventually be (A) extensively trained, (B) moderately trained, and (C) minimally trained for data from scanning sessions 2 (after approximately 2 weeks of training; circles), 3 (after approximately 4 weeks of training; squares), and 4 (after approximately 6 weeks of training; diamonds). The darkness of data points indicates scanning session; darker colors indicate earlier scans, so the darkest colors indicate scan 1 and the lightest ones indicate scan 4. Figure 3: Geometrical Core-Periphery Organization Over 42 Days. Geometrical core scores for each brain region defined by the trial blocks in which participants practiced sequences that would eventually be (A) extensively trained, (B) moderately trained, and (C) minimally trained for data from scanning sessions 1 (day 1; black circles), 2 (after approximately 2 weeks of training; dark gray squares), 3 (after approximately 4 weeks of training; gray diamonds), and 4 (after approximately 6 weeks of training; light gray stars). We have averaged the geometrical core scores over blocks and over 20 participants. The order of brain regions is identical for all 3 panels (A-C ), and we chose this order by ranking regions from high to low geometrical core scores from the EXT blocks on scanning session 1 (on day 1 of the experiment). 3 extensive training B moderate training C minimal training core bulk periphery 40 80 brain region 120 0 40 80 brain region 120 0 40 80 brain region 120 scan 2 scan 3 scan 4 0.18 0.14 0.1 0.06 0 A flexibility minimal training 50 brain region 100 0.5 0.4 0.3 0.2 0.1 C geometrical core score 0 0 moderate training 50 brain region 100 0.5 0.4 0.3 0.2 0.1 B geometrical core score 0 0 scan 1 scan 2 scan 3 scan 4 extensive training 50 brain region 100 0.5 0.4 0.3 0.2 0.1 A geometrical core score 0 0 j over participants and EXT blocks on day 1 of the experiment. We determined the community structure of this mean-coherence matrix by optimizing the single-layer modularity quality function [1, 2, 3, 4, 5]: kikj 2m(cid:21) δ(gi, gj) , Qsingle−layer =Xij (cid:20) ¯Aij − (1) where node i is assigned to community gi, node j is assigned to community gj, the Kronecker delta δ(gi, gj) = 1 if gi = gj and it equals 0 otherwise, ki is the strength of node i, and m is the mean strength of all nodes in the network. After optimizing this single-layer quality function 100 times, we constructed a representative partition [6] from the set of 100 partitions. (Each partition arises from a single optimization.) One community in this representative partition, which we show in Fig. 4A, appears to have high connectivity to the other two communities: nodes in this first community have edges with strong weights to nodes in the other two communities. This indicates a high coherence in the BOLD time series, and this behavior is consistent with the behavior expected from a network "core". A second community in this representative partition appears to have low connectivity to the other two communities: nodes in this community have edges with small weights that connect to nodes in the other two communities. This indicates a low coherence in the BOLD time series, and this behavior is consistent with the behavior expected from a "periphery". It is important to note that we observed this relationship between temporal core-periphery organization and community structure in networks encoded by mean matrices. However, networks encoded by mean ma- trices constructed by averaging correlation-based matrices often do not adequately represent the topological or geometrical structure of the ensemble of individual networks from which they are derived [7]. We therefore test for a relationship between the temporal core-periphery organization and community structure in the ensemble of networks extracted from individual participants. A division of the brain into temporal core, bulk, and peripheral regions gives a partition of the functional brain network. We label this partition using the Greek letter ν, and we use the z-score of the Rand coefficient [8] to test for similarities between this partition and algorithmic partitions, which we label using η, into communities (based on optimization of multilayer modularity) for each participant, block, and optimization. For each pair of partitions ν and η, we calculate the Rand z-score in terms of the total number of node pairs M in the network, the number of pairs Mν that are in the same community in partition ν but not in the partition η, the number of pairs Mη that are in the same community in partition η but not in ν, and the number of node pairs wνη that are assigned to the same community in both partition ν and partition η. The z-score of the Rand coefficient allows one to compare partitions η and ν, and it is given by the formula zνη = 1 σwνη wνη − MνMη M , (2) where σwνη is the standard deviation of wνη. Let the mean partition similarity zi denote the mean value of zνη over all partitions η (i.e., for all blocks and all optimizations) for participant i. As we show in Fig. 4B-D, we find that communities identified by the optimization of the multilayer mod- ularity quality function (see the "Materials and Methods" section in the main manuscript) have significant overlap with the division into temporal core, bulk, and periphery during early learning. The mean values of zi over participants indicate that there is a significant similarity between the partitions into modules and the partitions into core, bulk, and periphery for networks representing functional connectivity during blocks of extensively, moderately, and minimally trained sequences on scanning day 1. This similarity between community structure and temporal core-periphery organization is also evident for blocks of moderately and minimally trained sequences practiced during later scanning sessions. These results underscore the fact that core-periphery organization can be consistent with community structure. Note, however, that there is no statistical similarity between partitions into core, bulk, and periphery and partitions into communities for later learning. (As shown in Fig. 4B-D, the z-scores for networks that represent the functional connectivity during extensive training in scans 2 -- 4, moderate training in scans 3 -- 4, and minimal training in scan 4 are not significantly greater than expected (i.e., under the null hypothesis of no difference between the partitions).) Together, this set of results suggests that the relationship between these two types of mesoscale organization can be altered by learning. 4 Figure 4: Relationship Between Temporal Core-Periphery Organization and Community Struc- ture. (A) Mean-coherence matrix over all EXT blocks from all participants on scanning day 1. The colored bars above the matrix indicate the 3 communities that we identified from the representative partition. Mean partition similarity z-score zi over all participants for blocks of (B) extensively, (C) moderately, and (D) minimally trained sequences for all 4 scanning sessions over the approximately 6 weeks of training. The horizontal gray lines in panels (B-D) indicate the zi value that corresponds to a right-tailed p-value of 0.05. 5 extensive training 2 3 scanning session minimal training 2 3 scanning session 4 4 1 1 012345678 012345678 B z−score mean squared coherence 0.8 0.7 0.6 0.5 0.4 0.3 0.2 .1 00 D z−score 4 20 40 60 80 100 brain regions moderate training 1 2 3 scanning session 20 40 60 80 brain regions A 100 C 012345678 z−score Mean Minimum Maximum Standard Error Days Between Scans 1 and 2 Between Scans 2 and 3 Between Scans 3 and 4 Practice Sessions Between Scans 1 and 2 Between Scans 2 and 3 Between Scans 3 and 4 Extensively Trained Trials Between Scans 1 and 2 Between Scans 2 and 3 Between Scans 3 and 4 Moderately Trained Trials Between Scans 1 and 2 Between Scans 2 and 3 Between Scans 3 and 4 Minimally Trained Trials Between Scans 1 and 2 Between Scans 2 and 3 Between Scans 3 and 4 12.00 12.45 12.10 9.70 9.75 10.05 620.80 624.00 643.20 97.00 97.50 100.50 9.70 9.75 10.05 9 10 9 8 4 7 512 256 448 80 40 70 8 4 7 14 14 22 10 14 13 640 896 832 100 140 130 10 14 13 0.34 0.29 0.63 0.14 0.44 0.32 9.40 28.57 20.48 1.46 4.46 3.20 0.14 0.44 0.32 Table 1: Experimental Details for Behavioral Data Acquired Between Scanning Sessions. We give the minimum, mean, maximum, and standard error of the mean over participants for the following variables: the number of days between scanning sessions; the number of practice sessions performed at home between scanning sessions; and the number of trials composed of extensively, moderately, and minimally trained sequences during home practice between scanning sessions. Methodological Considerations Experimental Factors Effect of Region Size Recent studies have noted that brain-region size can affect estimates of hard-wired connectivity strength used in constructing structural connectomes [11, 12]. Although the present work is concerned with functional connectomes, it is nevertheless relevant to consider whether or not region size could be a driving effect of the observed core-periphery organization. Importantly, we observe no significant correlation between region size and flexibility (see Fig. 5), which suggests that region size is not driving the reported results. Effect of Block Design Another important factor is the underlying experimental block design and its effect on the correlation structure between brain regions in a single time window (i.e., in a single layer in the multilayer formalism). Two brain regions, such as motor cortex (M1) and supplementary motor area (SMA), might be active during the trial but quiet during the inter-trial interval (ITI). This would lead to a characteristic on-off activity pattern that is highly correlated with all other regions that also turn on with the task and off during the ITI. The frequency of this task-related activity (one on-off cycle per trial, where each trial is of length 4 -- 6 TRs) is included in our frequency band of interest (wavelet scale two, whose frequency range is 0.06 -- 0.12 Hz), and it therefore likely plays a role in the observed correlation patterns between brain regions in a single time window. Note, however, that our investigations of dynamic network structure -- namely, our computations of flex- ibility of community allegiance -- probe functional connectivity dynamics at much larger time scales, and the associated frequencies are an order of magnitude smaller. They lie in the range 0.0083 -- 0.012 Hz, as there 6 .= Figure 5: Region Size is Uncorrelated with Flexibility. (A) Scatter plot of the size of the brain region in voxels (averaged over participants) versus the flexibility of the EXT multilayer networks, which we averaged over the 100 multilayer modularity optimizations and the 20 participants. Data points indicate −0.009, and brain regions. The line indicates the best linear fit. Its Pearson correlation coefficient is r the associated p-value is p 0.92. (B) Box plot over the 20 participants of the squared Pearson correlation coefficient r2 between the participant-specific region size in voxels and the participant-specific flexibility averaged over the 100 multilayer modularity optimizations. .= is one time window every 40 -- 60 TRs. At these longer time scales, we can probe the effects of both early learning and extended learning independently of block-design effects. Specificity of Dynamic Network Organization as a Predictor of Learning An important consideration is whether there exist (arguably) simpler properties of brain function than flexibility that could be used to predict learning. We find that the power of activity, the mean connectivity strength, and parameter estimates from a general linear model (GLM) provide less predictive power than flexibility. Measures of Activity and Connectivity. It is far beyond the scope of this study to perform exhaustive computations using all possible measures of brain-region activity, so we focus on two common diagnostics. One is based on functional connectivity, and the other is based on brain activity. To estimate the strength of functional connectivity, we calculated the mean pairwise coherence between regional wavelet scale-two time series constructed from the BOLD signal, where we took the mean over all possible pairs of regions and all EXT experimental blocks extracted from scans on day 1 for a given subject. To estimate the strength of activity, we calculated the mean signal power of the regional wavelet scale-two time series constructed from the BOLD signal, where we took the mean over all regions and all EXT experimental blocks extracted from scans on day 1 for a given subject. We estimate the power Pw2 of the wavelet scale-two time series as the square of the time series normalized by its length: Pw2 =Xt w2(t)2 T , (3) where T is the length of the time series [13, 14]. .= .= .= .= We found that neither mean pairwise coherence nor mean power of regional activity measured during the first scanning session could be used to predict learning during the subsequent 10 home training sessions. For the mean pairwise coherence, we obtained a Pearson correlation of r 0.987. For the mean power of brain-region activity, we obtained r 0.354. These results indicate that a prediction similar to that made using the flexibility is not possible using the (arguably) simpler properties of the mean pairwise coherence or the mean power of regional brain activity. They also suggest that the dynamic pattern of coherent functional brain activity is more predictive than means of such activity patterns. −0.003 and a p-value of p −0.218 and p 7 0.04 0.03 0.02 0.01 0 B Pearson r2 500 0 region size (number of voxels) 1000 1500 2000 0.16 0.14 0.12 0.1 0.08 0.06 0.04 A flexibility Parameter Estimates for a General Linear Model. We determined relative differences in the BOLD signal by using a GLM approach for event-related functional data [15, 16]. For each participant, we con- structed a single design matrix for event-related fMRI by specifying the onset time and duration of all stimulus events from each scanning session (i.e., the pre-training session and the 3 test sessions). We found estimations of changes in the BOLD signal related to experimental conditions by using the design matrix with the GLM. We modeled the duration of each sequence trial as the time elapsed to produce the entire sequence; in other words, we calculated the movement time (MT), which is a direct measure of the time spent on a task and leads to accurate modeling of BOLD signals using the GLM [17]. Separate stimulus vectors indicate each sequence exposure type (EXT, MOD, and MIN) for each scanning session. We took potential differences in brain activity due to rate of movement into account by using the MT for each trial as the modeled duration for the corresponding event. We convolved events using the canonical hemodynamic response function and temporal derivative. Using the canonical hemodynamic response function (HRF) and its temporal derivative -- we use the implementation in the Statistical Parametric Mapping Toobox (SPM8) [18] -- we then modeled the events that were specified in the stimulus vectors. From this procedure, we obtained a pair of beta images for each event type. These images correspond to estimates of the HRF and its temporal derivative. Using freely available software [18], we then combined the corresponding beta image pairs for each event type (HRF and its temporal derivative) at the voxel level to form a magnitude image [19] H = sign( B1) +q( B1 + B2) , (4) where H is called the "combined amplitude" of the estimation of the BOLD signal using the HRF ( B1) and its temporal derivative ( B2). 1 This yielded separate magnitude images for each sequence exposure type (EXT, MOD, and MIN) and session. We then calculated the mean region-based magnitude for each exposure type and session using regions derived from each subject's grey matter-constrained Harvard-Oxford (HO) atlas. .= −0.10 and the p-value is p .= We did not find a significant correlation between the mean parameter estimates averaged over brain regions for the EXT trials in scanning session 1 and learning of the EXT sequences over the subsequent approximately 10 home training sessions. The Pearson correlation is r 0.65. Subject State-Dependence of Dynamic Network Organization Our finding that temporal core-periphery organization predicts the rate of learning across individuals is compelling evidence that the relationship between geometrical and temporal core-periphery organization is related to learning. Nevertheless, it is important to ask whether changes in dynamic community structure and associated mesoscale network organization are related to tasks or to changes in subjects' physiological state over the course of longitudinal imaging [20]. It is clear from studies of behavior, peripheral physiology, and fMRI that subjects can have high levels of anxiety or stress (particularly during their first exposure to MRI) [21]. To address this issue, we describe additional evidence that supports our conclusions that the reported changes in dynamic community structure with learning are indeed related to motor tasks. First, we note that we observed temporal and geometrical core-periphery organization consistently over all 4 scanning sessions. In Fig. 2 of the present document, we show that the anatomical identity of nodes in the temporal core, bulk, and periphery are consistent over scanning sessions. In Fig. 3 of this document, we show that the anatomical identity of nodes in the geometrical core and periphery are also consistent over scanning sessions. Moreover, Fig. 6 in the main manuscript shows that we observe the relationship between temporal and geometrical core-periphery organization consistently across scanning sessions. Second, we assume that the effects of a subject's mental and physiological state (e.g., anxiety) are greatest during the first imaging session [22]. If this is indeed the case, then there could be significant changes of network organization between scans 1 (higher anxiety) and 2 (lower anxiety) that might lead to a spurious interpretation of changes in core-periphery organization. To examine this possibility, we test whether the changes in dynamic community structure and core-periphery organization with learning are robust to the removal of scan 1. Importantly, the trends in Figs. 2 and 5 in the main manuscript remain present if we only examine scans 2 -- 4. We use data from scan 1 for the three box plots located at the point in the horizontal 1In this equation, we use the hat notation to indicate that these values are estimated (rather than directly measured) from a general linear model for a response variable (such as regional cerebral blood) at each voxel in a given participant [16]. 8 axis at which the number of trials is equal to 50. (This is the leftmost point of each panel.) See Table 1 in the main manuscript. The 9 box plots located at points on the horizontal axis at which the number of trials is greater than 50 use data from scans 2 -- 4. Therefore, when we examine only scans 2 -- 4, we still observe a decrease in maximum modularity, an increase in the number of communities, an increase in flexibility, and a decrease in the variance of the geometrical core score with learning. .= .= .= .= −0.52 (so p .= 0.615, p .= 0.00137) and a non-significant effect of scanning session (F (3, 57) Finally, task-related fMRI BOLD activation magnitude in core, bulk, and peripheral brain regions are not altered significantly across scanning sessions. We employed a repeated-measures analysis of variance (ANOVA) on the training-depth-averaged GLM parameter estimates [23]. We treated core, bulk, and periph- ery designations as categorical factors, and we treated scanning session as a repeated measure. We found a sig- nificant main effect (i.e., single-factor effect) of core, bulk, and periphery (an F-statistic [23] of F (2, 38) 7.88 and a p-value of p 0.584). These results suggest that a systematic change in the hemodynamic response function across scanning ses- sions is unlikely to be responsible for the observed learning-related changes in dynamic community structure. Furthermore, we observe that mean GLM parameter estimates in core, bulk, and peripheral brain regions are not correlated significantly with the reported changes in core-periphery structure that accompany learn- ing. The Pearson correlation coefficient between parameter estimates and the variance of the geometrical core score for nodes in the temporal core is r 0.52), for nodes in the temporal bulk is r 0.08). These results provide further evidence that BOLD activation magnitude and dynamic community structure provide distinct insights. 0.86), and for nodes in the temporal periphery is r 0.20 (which gives a p-value of p .= −0.05 (so p .= .= .= Temporal Core-Periphery Organization and Task-Related Activations One of the strengths of our approach is that we examine the organization of whole-brain functional connec- tivity and thereby remain sensitive to a wide variety of learning-related changes in the brain that could not be identified using a traditional GLM analysis. Nevertheless, it is useful to explore the relationship between dynamic community structure and task-related activations. In Fig. 6, we show that regions in the temporal core tend to be regions with strong task-related activations, as evinced by high (and positive) values of mean GLM parameter estimates. Conversely, regions in the temporal bulk and periphery tend to lack strong task-related activations, as evinced by low (and negative) values of mean GLM parameter estimates. These results are consistent with our interpretation that the temporal core consists of a small set of regions that are required to perform a given task and that the temporal periphery consists of a set of regions that are associated more peripherally with the task and which are activated in a transient manner. Dynamic Community Detection In the multilayer modularity quality function (see the "Materials and Methods" section of the main manuscript), we need to choose values for two parameters [6]: a structural resolution parameter γ and a temporal resolution parameter ω. We now examine the effects of these choices on our results. Effect of Structural Resolution Parameter In the main manuscript, we used a structural resolution parameter value of γ = 1, which is the most common choice when optimizing the single-layer and multilayer modularity quality functions [4, 5, 24]. In this case, A − γP = A − P, and one is simply subtracting the optimization null model P from the adjacency tensor A. One can decrease γ to access community structure at smaller spatial scales (i.e., to examine smaller communities) or increase it to access community structure at larger spatial scales (i.e., to examine larger communities). By examining network diagnostics over a range of γ values, we explore the spatial specificity of our results. The mean number of communities in the partitions that we obtained by optimizing multilayer modularity Q varies from the minimum (1) to the maximum (112) possible value for γ approximately in the interval [0.8, 2.5] (see Fig. 7A). We investigate this transition in greater detail in Figs. 7C,D. Near the value γ = 1, the number of regions in the bulk dips to about 65, whereas the number of regions in the core and periphery rise to about 20 and 25, respectively. Observe the dip of the bulk curve and bumps of the core and periphery curves in Fig. 7D. These features occur for γ approximately in the interval [0.88, 1.22], which corresponds 9 Figure 6: Temporal Core-Periphery Organization and Task-Related Activations. Mean GLM parameter estimates for the temporal core (cyan; circles), bulk (gold; squares), and periphery (maroon; diamonds) of dynamic networks defined by the trial blocks in which participants practiced sequences that would eventually be (A) extensively trained, (B) moderately trained, and (C) minimally trained for data from scanning sessions 1 (first day of training), 2 (after approximately 2 weeks of training), 3 (after approximately 4 weeks of training), and 4 (after approximately 6 weeks of training). Figure 7: (Supplementary Material) Effect of Structural Resolution Parameter. (A,B) Number of communities and (C,D) number of regions in the temporal core (cyan; circles), temporal bulk (gold; squares), and temporal periphery (maroon; diamonds) as a function of the structural resolution parameter γ, where we considered (A,C) γ ∈ [0.2, 5] in increments of ∆γ = 0.2 and (B,D) γ ∈ [0.8, 1.8] in increments of ∆γ = 0.01. We averaged the values in panels (A) and (B) over 100 multilayer modularity optimizations and over the 20 participants. 10 minimal training 2 3 scanning session 4 1 0.5 0 −0.5 C mean parameter estimates 4 −1 1 moderate training 1 0.5 0 −0.5 B mean parameter estimates extensive training core bulk periphery 2 3 scanning session 4 −1 1 2 3 scanning session 1 0.5 0 −0.5 A mean parameter estimates −1 1 100 80 60 40 20 D number of regions core bulk periphery 80 80 60 60 40 40 20 20 0 0 1.8 0.8 0 5 structural resolution parameter structural resolution parameter 1.2 2 1.4 3 1.6 4 1 1 100 100 B number of communities number of regions 100 80 60 40 20 0 0 5 structural resolution parameter 1 2 3 4 number of communitiesA C 100 100 80 80 60 60 40 40 20 20 number of communities number of regions 0 0 0 5 1.8 0.8 structural resolution parameter structural resolution parameter 2 1.2 3 1.4 4 1.6 1 1 0 0.8 1.8 structural resolution parameter 1.6 1.4 1.2 1 to partitions that are composed of between approximately 3 and approximately 20 communities (with an associated mean community size of between approximately 6 and approximately 37 brain regions; see Fig.7B). This supports our claim that the temporal core-periphery structure that we examine in this study is a genuine mesoscale feature of coherent brain dynamics. Effect of Temporal Resolution Parameter In the main manuscript, we used a temporal resolution parameter value of ω = 1. The value ω = 1 ensures that the inter -layer coupling is equal to the maximum possible value of the intra-layer coupling, which we compute from the magnitude-squared coherence (which is constrained to lie in the interval [0, 1]). It is important to examine the robustness of results for different values of this parameter, and investigating dynamic network structure at other values of ω can also provide additional insights [6]. For example, one can decrease ω to encourage greater variability in community assignments of nodes across individual layers (i.e., across time in temporal networks) or increase it to encourage such community assignments to be more similar across layers. Recall that each node in the temporal multilayer network represents a single brain region at a specified time, and different nodes that represent the same brain region at different times become more likely to be assigned to the same multilayer community as ω is increased. By examining network diagnostics over a range of ω values, we can quantify the robustness of our results to differing amounts of temporal variation in community structure. We varied ω from 0.1 to 2 in increments of ∆ω = 0.1. As expected, we find that the number of communities identified in the optimization of the multilayer modularity quality function decreases as ω is increased (see Fig. 8A). This is consistent with the fact that greater variation of community assignments across time is possible for smaller values of ω. Variation between community assignments of nodes in individual layers can occur in two ways: (1) a small number of regions change community membership from one layer to the next, but the majority of regions retain their community membership; or (2) entire communities lose their identities (via fragmentation, extinction, union, and/or recombination), such that the algorithm identifies either the "death" of a community that was present in the previous layer but is not present in the current layer or the "birth" of a community that was not present in the previous layer but is present in the current layer. For each value of ω, we examined the robustness of our division of brain regions into a temporal core, a temporal bulk, and a temporal periphery using the same procedure that we employed for ω = 1. Namely, we defined a temporal core and temporal periphery as those brain regions that were composed, respectively, of the brain regions below and above the 95% confidence interval of the nodal null model. In Fig. 8B, we report the number of regions in each group as a function of ω. Interestingly, the number of brain regions that we identified as part of the temporal core varied little over the examined range of ω values; it remained at approximately 17.0±1.1. In fact, 15 of the 17 regions that we identified as part of the temporal core at ω = 1 were also identified as part of the temporal core at all other values of ω that we examined. The number of regions in the temporal bulk and temporal periphery varied more (with values of approximately 75.6 ± 7.4 for the bulk and approximately 19.4± 6.8 for the periphery), which suggests that the separation between the temporal bulk and temporal periphery is less drastic than that between temporal core and temporal bulk. Indeed, the mean flexibility of the core is less similar to the mean flexibility of the bulk than is the latter to the mean flexibility of the periphery. See Fig. 3 of the main manuscript and Figs. 1 and 2 of this supplement. References [1] Newman MEJ, Girvan M (2004) Finding and evaluating community structure in networks. Phys Rev E 69: 026113. [2] Newman MEJ (2004) Fast algorithm for detecting community structure in networks. Phys Rev E 69: 066133. [3] Newman MEJ (2006) Modularity and community structure in networks. Proc Natl Acad Sci USA 103: 8577 -- 8582. 11 Figure 8: Effect of Temporal Resolution Parameter. (A) Number of communities averaged over 100 multilayer modularity optimizations and over 20 participants as a function of the temporal resolution parameter ω. (B) Number of regions that we identified as part of the temporal core (cyan; circles), temporal bulk (gold; squares), and temporal periphery (maroon; diamonds) as we vary ω from 0.1 to 2 in increments of ∆ω = 0.1. [4] Porter MA, Onnela JP, Mucha PJ (2009) Communities in networks. Not Amer Math Soc 56: 1082 -- 1097, 1164 -- 1166. [5] Fortunato S (2010) Community detection in graphs. Phys Rep 486: 75 -- 174. [6] Bassett DS, Porter MA, Wymbs NF, Grafton ST, Carlson JM, et al. (2012) Robust detection of dynamic community structure in networks. Chaos 23: 013142. [7] Simpson SL, Moussa MN, Laurienti PJ (2012) An exponential random graph modeling approach to creating group-based representative whole-brain connectivity networks. NeuroImage 60: 1117 -- 1126. [8] Traud AL, Kelsic ED, Mucha PJ, Porter MA (2011) Comparing community structure to characteristics in online collegiate social networks. SIAM Rev 53: 526 -- 543. [9] Smith SM, Jenkinson M, Woolrich MW, Beckmann CF, Behrens TEJ, et al. (2004) Advances in func- tional and structural MR image analysis and implementation as FSL. NeuroImage 23: 208 -- 219. [10] Woolrich MW, Jbabdi W, Patenaude B, Chappell M, Makni S, et al. (2009) Bayesian analysis of neuroimaging data in FSL. NeuroImage 45: S173 -- S186. [11] Hagmann P, Cammoun L, Gigandet X, Meuli R, Honey CJ, et al. (2008) Mapping the structural core of human cerebral cortex. PLoS Biol 6: e159. [12] Bassett DS, Brown JA, Deshpande V, Carlson JM, Grafton ST (2011) Conserved and variable archi- tecture of human white matter connectivity. NeuroImage 54: 1262 -- 1279. [13] MATLAB (2012) Measuring signal power. MATLAB's Help Menu . [14] Smith SW (1997) Digital Signal Processing: A Guide for Engineers and Scientists. California Technical Pub. [15] Lazar N (2010) The Statistical Analysis of Functional MRI Data. Springer-Verlag. [16] Friston KJ, Holmes AP, Worsley KJ, Poline JP, Frith CD, et al. (1994) Statistical parametric maps in functional imaging: A general linear approach. Human Brain Mapping 2: 189 -- 210. 12 core bulk periphery B 100 80 60 40 20 number of regions 0 0 2 temporal resolution parameter 0.5 1.5 1 0 2 temporal resolution parameter 1.5 0.5 1 234567 number of communitiesA [17] Grinband J, Wager TD, Lindquist M, Ferrera VP, Hirsch J (2008) Detection of time-varying signals in event-related fMRI designs. NeuroImage 43: 509 -- 520. [18] Steffener J, Tabert M, Reuben A, Stern Y (2010) Investigating hemodynamic response variability at the group level using basis functions. NeuroImage 49: 2113 -- 2122. [19] Calhoun VD, Adali T, Pekar JJ (2004) A method for comparing group fMRI data using independent component analysis: application to visual, motor and visuomotor tasks. Magn Reson Imaging 22: 1181 -- 1191. [20] Johnstone T, Somerville LH, Alexander AL, Oakes TR, Davidson RJ, et al. (2005) Stability of amygdala BOLD response to fearful faces over multiple scan sessions. NeuroImage 25: 1112 -- 1123. [21] Lueken U, Muehlhan M, Evens R, Wittchen HU, Kirschbaum C (2012) Within and between session changes in subjective and neuroendocrine stress parameters during magnetic resonance imaging: A controlled scanner training study. Psychoneuroendocrinology 37: 1299 -- 1308. [22] Chapman HA, Bernier D, Rusak B (2010) MRI-related anxiety levels change within and between re- peated scanning sessions. Psychiatry Res 182: 160 -- 164. [23] Agresti A, Franklin CA (2007) Statistics: The Art and Science of Learning From Data. Prentice Hall. [24] Reichardt J, Bornholdt S (2006) Statistical mechanics of community detection. Phys Rev E 74: 016110. 13 Minimum Mean Maximum Standard Error Extensively Trained Blocks During Scan 1 During Scan 2 During Scan 3 During Scan 4 Moderately Trained Blocks During Scan 1 During Scan 2 During Scan 3 During Scan 4 Minimally Trained Blocks During Scan 1 During Scan 2 During Scan 3 During Scan 4 Length of Extensively Trained Blocks During Scan 1 During Scan 2 During Scan 3 During Scan 4 Length of Moderately Trained Blocks During Scan 1 During Scan 2 During Scan 3 During Scan 4 Length of Minimally Trained Blocks During Scan 1 During Scan 2 During Scan 3 During Scan 4 6.00 10.00 10.00 8.00 5.00 10.00 10.00 8.00 7.00 10.00 10.00 8.00 52.50 35.50 35.40 34.60 50.80 39.70 37.60 37.60 52.10 44.10 42.50 39.70 9.70 10.00 10.00 9.90 9.70 10.00 10.00 9.90 9.80 10.00 10.00 9.90 61.94 42.36 40.79 40.30 61.67 47.56 45.07 43.83 61.19 50.02 47.37 45.79 10.00 10.00 10.00 10.00 11.00 10.00 10.00 10.00 11.00 10.00 10.00 10.00 72.20 45.90 45.50 45.70 72.60 57.20 52.80 50.60 70.60 57.70 54.50 54.10 0.21 0.00 0.00 0.10 0.27 0.00 0.00 0.10 0.18 0.00 0.00 0.10 1.34 0.72 0.77 0.87 1.26 0.80 0.67 0.79 1.29 0.73 0.71 0.70 Table 2: Experimental Details for Brain Imaging Data Acquired During Scanning Sessions. In the top three rows, we give the mean, minimum, maximum, and standard error over participants for the number of blocks composed of extensively, moderately, and minimally trained sequences during scanning sessions. In the bottom three rows, we give (in TRs) the mean, minimum, maximum, and standard error of the length over blocks composed of extensively, moderately, and minimally trained sequences during scanning sessions. 14 Region Name Frontal pole Insular cortex Superior frontal gyrus Middle frontal gyrus Affiliation Region Name B(R) B(L) Cingulate gyrus, anterior B(R) B(L) Cingulate gyrus, posterior B(R) B(L) Precuneus cortex B(R) B(L) Cuneus cortex Inferior frontal gyrus, pars triangularis B(R) P(L) Orbital frontal cortex Affiliation B(R) B(L) P(R) B(L) B(R) B(L) C(R) C(L) B(R) B(L) Inferior frontal gyrus, pars opercularis B(R) B(L) Parahippocampal gyrus, anterior B(R) B(L) Precentral gyrus Temporal pole C(R) C(L) Parahippocampal gyrus, posterior B(R) P(L) B(R) B(L) Lingual gyrus C(R) C(L) Superior temporal gyrus, anterior B(R) B(L) Temporal fusiform cortex, anterior P(R) B(L) Superior temporal gyrus, posterior B(R) B(L) Temporal fusiform cortex, posterior P(R) P(L) Middle temporal gyrus, anterior B(R) B(L) Temporal occipital fusiform cortex P(R) P(L) Middle temporal gyrus, posterior B(R) B(L) Occipital fusiform gyrus Middle temporal gyrus, temporooccipital P(R) P(L) Frontal operculum cortex Inferior temporal gyrus, anterior B(R) B(L) Central opercular cortex Inferior temporal gyrus, posterior B(R) B(L) Parietal operculum cortex Inferior temporal gyrus, temporooccipital B(R) B(L) Planum polare Postcentral gyrus P(R) C(L) Heschl's gyrus Superior parietal lobule B(R) C(L) Planum temporale Supramarginal gyrus, anterior B(R) C(L) Supercalcarine cortex Supramarginal gyrus, posterior B(R) P(L) Occipital pole Angular gyrus P(R) B(L) Caudate Lateral occipital cortex, superior P(R) P(L) Putamen Lateral occipital cortex, inferior P(R) B(L) Globus pallidus Intracalcarine cortex Frontal medial cortex C(R) C(L) Thalamus B(R) B(L) Nucleus Accumbens Supplemental motor area C(R) C(L) Parahippocampal gyrus Subcallosal cortex Paracingulate gyrus B(R) B(L) Hippocampus B(R) B(L) Brainstem P(R) P(L) B(R) P(L) B(R) B(L) P(R) B(L) C(R) B(L) C(R) C(L) B(R) B(L) C(R) C(L) C(R) B(L) P(R) B(L) P(R) B(L) P(R) B(L) P(R) P(L) B(R) B(L) B(R) B(L) B(R) B(L) B(R) B(L) Table 3: Brain regions in the Harvard-Oxford (HO) Cortical and Subcortical Parcellation Scheme provided by FSL [9, 10] and their affiliation to the temporal core (C; cyan), bulk (B; gold), and periphery (P; maroon) for both left (L) and right (R) hemispheres. 15
1712.00462
1
1712
2017-12-01T19:09:32
Functional brain networks reveal the existence of cognitive reserve and the interplay between network topology and dynamics
[ "q-bio.NC" ]
We investigated how the organization of functional brain networks was related to cognitive reserve (CR) during a memory task in healthy aging. We obtained the magnetoencephalographic functional networks of 20 elders with a high or low CR level to analyse the differences at network features. We reported a negative correlation between synchronization of the whole network and CR, and observed differences both at the node and at the network level in: the average shortest path and the network outreach. Individuals with high CR required functional networks with lower links to successfully carry out the memory task. These results may indicate that those individuals with low CR level exhibited a dual pattern of compensation and network impairment, since their functioning was more energetically costly to perform the task as the high CR group. Additionally, we evaluated how the dynamical properties of the different brain regions were correlated to the network parameters obtaining that entropy was positively correlated with the strength and clustering coefficient, while complexity behaved conversely. Consequently, highly connected nodes of the functional networks showed a more stochastic and less complex signal. We consider that network approach may be a relevant tool to better understand brain functioning in aging.
q-bio.NC
q-bio
Functional brain networks reveal the existence of cognitive reserve and the interplay between network topology and dynamics Johann H. Martínez *, María E. López *, Pedro Ariza, Mario Chavez, José A. Pineda- Pardo, David López-Sanz, Pedro Gil, Fernando Maestú and Javier M. Buldú. 1 INSERM, Institut du Cerveau et de la Moelle Epinière (ICM). Hôpital Pitié Salpétrière, Paris, France 2 Laboratory of Biological Networks, Center for Biomedical Technology (CTB), Madrid, Spain 3 Department of Basic Psychology II, Complutense University of Madrid, Madrid, Spain 4 7CNRS-UMR 7225, Hôpital Pitié-Salpetrière, Paris, France. 5 Centro Integral de Neurociencias AC (CINAC), HM Puerta del Sur, Madrid, Spain 6 CEU San Pablo University, Madrid, Spain. 7Laboratory of Cognitive and Computational Neuroscience (UCM-UPM), Center for Biomedical Technology (CTB), Madrid, Spain 8 Institute of Sanitary Investigation [IdISSC], San Carlos University Hospital, Madrid, Spain 9 Geriatrics Department, San Carlos University Hospital, Madrid, Spain. 10 Complex Systems Group, Universidad Rey Juan Carlos, Madrid, Spain * These authors have contributed equally and must be considered as the first authors ABSTRACT We investigated how the organization of functional brain networks was related to cognitive reserve (CR) during a memory task in healthy aging. We obtained the magnetoencephalographic functional networks of 20 elders with a high or low CR level to analyse the differences at network features. We reported a negative correlation between synchronization of the whole network and CR, and observed differences both at the node and at the network level in: the average shortest path and the network outreach. Individuals with high CR required functional networks with lower links to successfully carry out the memory task. These results may indicate that those individuals with low CR level exhibited a dual pattern of compensation and network impairment, since their functioning was more energetically costly to perform the task as the high CR group. Additionally, we evaluated how the dynamical properties of the different brain regions were correlated to the network parameters obtaining that entropy was positively correlated with the strength and clustering coefficient, while complexity behaved conversely. Consequently, highly connected nodes of the functional networks showed a more stochastic and less complex signal. We consider that network approach may be a relevant tool to better understand brain functioning in aging. Keywords: Magnetoencephalography, Cognitive Reserve, Healthy aging, Brain networks, Efficiency. INTRODUCTION In recent years, there is an increased interest in studying lifestyle factors that may help to achieve a successful aging. These factors, such as educational level, leisure activities or intelligence quotient (IQ) have been related to the concept of reserve, which emerges from the absence of a direct relationship between the severity of brain pathology and its clinical manifestations 1. Two non-mutually exclusive models have been proposed to address these counterintuitive observations 1: a passive or brain reserve (BR) model, which considers that the amount of neural substrate (e.g. the brain size or neuron quantity) determines the threshold beyond which clinical and functional deficits emerge 2,3; and an active or cognitive reserve (CR) model, that refers to the ability to use the existent resources as flexible and efficiently as possible during the execution of cognitive tasks (neural reserve) or when coping with brain pathology (neural compensation) 4. Healthy aging and dementia, especially Alzheimer's disease (AD), have been considered optimal models to understand the role of reserve 4–6. Among lifestyle variables, the educational level (alone or in combination with occupational attainment) is one of the most studied and important CR proxies. In fact, some studies have considered the educational level as a "protective" factor by reducing the incidence of dementia 5, and its pathological consequences in AD patients 7,8. However, recent studies did not find this effect 9,10, suggesting that CR actually may have a "masking" effect. One of the new approaches to explore how brain functioning may be modulated by CR is functional network organization (based on functional connectivity; FC), although there is still scarcity of studies focused on this field. Solé-Padullés et al. 11 described that healthy elders with higher CR showed reduced functional magnetic resonance imaging (fMRI) activation during cognitive processing. In the same line, a previous study from our group 12 observed that healthy elders with a high CR score in comparison with those with low CR score, exhibited less magnetoencephalographic (MEG) functional connectivity during the performance of a memory task. These results may suggest that higher CR is related to more effective use of cerebral resources although this hypothesis has not been tested yet. In addition, FC may be used to explore brain's organization in terms of structural and functional networks, by the application of different methodologies coming from Network Science 13. This kind of analysis can introduce a different perspective to better understand how CR modules functional resilience in healthy aging. Recently, Yoo et al. 14 and Marques et al. 15 have proposed the analysis of the topology of the associated functional networks as an alternative way of quantifying CR. Interestingly, functional brain networks obtained during resting state show a positive correlation between CR and network efficiency 16,17. Network efficiency, first introduced by Latora et al. 18 is a metric evaluating the inverse of the topological distance between nodes of a network. In the context of functional brain networks, the topological distance between two brain regions is normally defined as the inverse of their level of coordination. In this way, nodes that are not functionally connected (i.e., with zero coordination) have infinite distance between them. However, information between not connected regions (at an infinite topological distance) can travel indirectly through intermediate nodes. Network efficiency measures how close or distant the nodes of a network are in terms of their topological distance using the functional connections through intermediate nodes 18.Under this framework, the positive correlation between CR and network efficiency during resting state relies on fMRI recordings, which report higher levels of synchronization between brain regions in individuals with higher CR15,16. However, during a cognitive process, the correlation between CR and synchronization turns to be negative, as shown in 11. With the aim of clarifying whether CR is related to the topology of the functional networks during a memory task, we are concerned about the biomagnetic network profiles of two groups of healthy elders with different CR levels. Specifically, we obtained the MEG recordings of both groups during the performance of a modified Stenberg's test 17 and we explored the topological properties of the corresponding functional networks. Among the diversity of network metrics that one may extract from the functional networks we focused on the most extended ones: the network strength, which is a measure of the average level of synchronization along the functional network 19; the weighted clustering coefficient, measuring the existence of triangles in the network and related to the local resilience 20; the eigenvector centrality a measure of node importance obtained from the eigenvector associated to the largest eigenvalue of the adjacency matrix 21; the average shortest path d(i), measuring the average number of steps to travel from one node to another through the shortest topological distance 22; the strength of the nearest neighbors snn(i), indicating what is the average strength (i.e., sum of the weights of a node) of the neighbors (i.e., nodes directly connected) of a given node 19; and the outreach o(i), a network metric obtained by multiplying the weight of the connections by their Euclidean distance 23 (see Supp. Info. for a detailed definition of all network metrics). We also quantified the dynamical properties of the brain regions in terms of entropy and complexity to compare both high and low CR groups. To the best of our knowledge, the present MEG network's approach, where the study of network metrics is combined with the dynamical properties of the nodes, has never been used within this field of research. Combining both approaches we have been able to detect differences between low and high CR individuals at the network level, showing that the network outreach and, consequently, the energetic expenses, is higher in individuals with low CR. At the same time, we show how the topological properties of the nodes (i.e., brain regions) are related to the complexity and entropy of their corresponding time series. MATERIALS AND METHODS Participants The sample consisted of 20 healthy elders recruited from the "Geriatric Unit of the University Hospital San Carlos" (Madrid, Spain). All of them were right-handed 24, native Spanish speakers, and between 65 and 85 years old. Regarding the neuropsychological assessment, all of them met the following criteria: 1) a score above 28 in MMSE 25; 2) no memory impairment as evidenced by delayed recall from Logical Memory II subtest of the Wechsler Memory Scale Revised (WMS- III-R; 26); and 3) normal daily living activities measured by the Spanish version of the Functional Assessment Scale 27. Exclusion criteria included: 1) a history of psychiatric or neurological disease; 2) psychoactive drugs consumption; and 3) severe sensory or comprehension deficits. The 20 participants did not differ in physical, cognitive and social activities during late-life, and were further subdivided into two subgroups according to their CR level. The CR index (CRI) was estimated for each individual by adding the following two categories: 1) formal educational level (from 1 –illiterate/ functional illiterate- to 5- superior studies-) and 2) occupational attainment (from 1 –housewife- to 5 -business management/research-). According to their CRI score, 12 participants were classified into the low CR group if their CRI score was between 1 and 5; or into the high CR group (8 subjects) if their CRI score was between 6 and 10. In addition, both groups did not differ in age and in their performance during the cognitive task (see table 1). Ethics statement: Methods were carried out in "accordance" with the approved guidelines. The investigation was approved by the local Ethics Committee (Madrid) and all participants signed a written informed consent before the MEG recordings. Age High- CR (n=8) Low- CR (n=12) P values 67.3±7.4 69.7±6.6 Educational Level 4.5±0.7 Occupation attainment 4.5±0.5 CRI score Cognitive activity Physical activity Social activity 9±0.6 4.2±2.3 2.4±1.5 1.6±1.2 2.4±0.7 1.6±0.5 3.7±1.1 2.8±1.4 2.7±1 1.7±1 Acc. Smaqe 114±10.7 106±14.1 p>0.05 p<0.05* p<0.05* p<0.05* p>0.05 p>0.05 p>0.05 p>0.05 Table 1. Mean values ± standard deviation for high and low CR groups. Significant p values are p<0.05 and marked with an asterisk (*). Both groups differ in their educational level, occupation attainment and CRI scores, but not in the other variables. Acc. Smaqe, accuracy in the modified Stenberg's memory task. MEG task A modified version of the Sternberg's letter probe task was employed 17 (see Figure 1). Each subject was asked to remember a single set of 5 letters (i.e. 'SMAQE'). After that, there were presented a series of single letters, one at a time (500ms in duration with a random ISI between 2 and 3 seconds). They were asked to make a match/non-match decision by pressing a button with their right hand if the presented letter was one of the five initially memorized. The experiment consisted of 250 letters, in which half were initially presented and the rest were distracters, meaning they were not in the initial 5 letter set. All of the participants completed a training session until demonstrating they had correctly memorized the initial letter set. The task was projected through a LCD video projector (SONY VPLX600E) located outside the shielded MEG room through a series of mirrors. The screen was suspended approximately 1 meter above the participant's face. The letters subtended 1, 8 and 3 degrees of horizontal and vertical visual angle respectively. Figure 1. Representation of the modified version of the Sternberg's letter probe task. In the encoding phase, participants are instructed to memorize 5 letters (i.e.: "SMAQE"). In the recognition phase, participants are instructed to make a match/non-match button-press to indicate that the presented letter matched any of the encoded ones. MEG recordings MEG data was recorded during the execution of the modified Sternberg's task with a 254 Hz sampling rate and a band pass of 0.5 to 50 Hz, using a 148 channel whole-head magnetometer (MAGNES 2500 WH, 4D Neuroimaging) placed in a magnetically shielded room. In order to reduce external noise we employed an algorithm using reference channels at a distance from the MEG sensors. After trial segmentation, artifacts were visually inspected by an expert and epochs containing muscular artifacts or eye movements were discarded for further analysis. Only hits were considered, removing false alarms, correct rejections and omissions. We randomly selected a set of 35 trials for each individual, since it was the lowest number of successful epochs for all subjects in the study. Each trial consisted of each 230 time points for each of the 148 channels. Trials are not necessary consecutive, however, this issue is not crucial for the analysis of the coordination between sensors since each trial is analyzed individually and averaged later. Functional connectivity analysis To calculate the connectivity networks for each subject we used Synchronization Likelihood algorithm (SL) 28 which is a nonlinear measure of the synchronized activity that has been proven to be a suitable quantifier for datasets obtained from magnetoencephalographic recordings 29. Specifically, the SL algorithm detects windows of repeated patterns within the time series of a channel A and, next, checks whether a channel B also shows a repeated pattern at the same time windows, no matter if it is the same or different to that observed in channel A. The range of values of SL is 0 ≤ SL ≤ 1, being 0 when the time series of both channels are uncorrelated, and 1 for maximal synchronization 28. SL was calculated for each pair of sensors since, in our study, sensors were the nodes of the functional networks. All pairs of SL are then included into a correlation matrix 𝑊{𝑤𝑖𝑗} where 𝑤𝑖𝑗 have values comprised between ~ 0.05 and ~ 0.5. We apply a linear normalization that leads to a probability matrix 𝑃{𝑝𝑖𝑗}, where the values of 𝑝𝑖𝑗 are obtained as 𝑝𝑖𝑗 = 𝑤𝑖𝑗 − min [𝑤𝑖𝑗] max[𝑤𝑖𝑗] − min [𝑤𝑖𝑗] . (1) In this way, the probability matrix 𝑃{𝑝𝑖𝑗}, whose values are within the interval [0,1], reflects the probability of the presence of a link between nodes (brain regions) i and j. Graph metrics We calculated a series of metrics related with the role of the nodes within the network: the strength s(i), the weighted clustering cw(i), the eigenvector centrality ec(i), the average shortest path d(i), the strength of the nearest neighbours snn(i) and the outreach o(i). We used the inverse of the values of the probability matrix 𝑃{𝑝𝑖𝑗} for obtaining the "topological distances" between nodes to quantify the shortest paths. Another parameters such as the within-module degree z(i) (also kwon as the z-score) and the participation coefficient p(i) 23,30 were also computed using the community affiliation vector Ccom(i) extracted from the classical partition of the brain into six lobes: left-frontal, right-frontal, central, left-temporal, right-temporal and occipital. A series of global network features such as the global efficiency Eg and the average shortest path (d) were also calculated. The node average (average along the same node for all subjects within the same group) of several metrics was computed to obtain the following mean values: 𝑠(𝑖), 𝑐𝑤(𝑖), 𝑒𝑣(𝑖), 𝑧(𝑖), 𝑑(𝑖), ), 𝑠𝑛𝑛(𝑖) and 𝑜(𝑖). Network averages of the preceding features were also appraised in order to compare both groups. Finally, we obtained the global efficiency 𝐸𝑔 of the networks from the harmonic mean of the inverse of the shortest paths. We constructed a set of randomized versions of the former functional networks, in order to evaluate to what extent the deviation of the network parameters between groups was a consequence of a topological reorganization or, on the contrary, just a matter of the number of links and their weights. With this purpose, we generated a group of 100 networks for each of the functional network. The randomization maintains the value of the links' weights by reshuffling the components of the weighted probability matrix 𝑃{𝑝𝑖𝑗} (see 23 for a similar normalization). In this way, we guaranteed that the average strength of the network was maintained. Next, we calculated the network parameters for each of the randomized versions and obtained an average value of each metric. Finally, we normalized all network parameters with the average of the set of surrogate matrices, i.e., for a parameter X its normalized value would be: 𝑋 = 𝑋 𝑋𝑟𝑎𝑛𝑑 (2) Note that this normalization allows focusing on changes related to the structure of the functional networks, since they exclude variations of the average weights of the networks (and their influence on the network metrics). Evaluation of the dynamical properties of the nodes In order to calculate the entropy and complexity brain signal, we used for each of the nodes in the network their corresponding time series comprising the 35 trials of the whole experiment. Since each trial has 230 time steps, we obtained times series of M = 8050 points for each node of each functional network. Next, we applied the methodology introduced by 31 to quantify the dynamical properties of nonlinear dynamical systems. We calculated the normalized permutation entropy of the signal (S) as described in 32 to capture the level of uncertainty of a signal. We calculated the complexity (H) of the signal for each subject following the procedure described in 33. This measure is effective to quantify the complexity of different dynamical systems 34. Statistical analysis We performed permutation tests as well as non-parametric ones for all statistical hypotheses of this work. The permutation test is based on a t-statistical distance between two samples. For robustness, we constructed a distribution of 5000 different p-values associated to the number of permutations in each hypothesis. In each distribution, we found the percentile associated to the original p-value, thus dividing this percentile by the number of permutations as a correction of the resampling. We considered the rejection of the null hypothesis when the aforementioned ratio is bellow or equal to α = 0.05 level of significance. Results of permutation tests are presented throughout the paper. Complimentary, we performed a nonparametric Mann-Whitney U-test (rank-sum test) to show how the results do not change. The latter results are shown in the Supplementary Information. RESULTS Micro-scale: Differences at the node level Firstly, we analysed those nodes with the highest eigenvector centrality in both groups (𝑒𝑣ℎ𝑖𝑔ℎ,𝑙𝑜𝑤). Figure 2 shows that both groups exhibited the most influencing nodes placed at the same cortical regions (i.e., at the occipital lobe). Nevertheless, we observed signs of a displacement of the node centrality toward the central and left-temporal lobes. Figure 2. Back view for eigenvector centrality (𝒆𝒗(𝒊)) for high (A) and low (B) CR groups. Circles highlight the ten nodes with the highest 𝒆𝒗(𝒊). Note that they are localized close to the occipital lobe in the low CR group (blue circles) and slightly shift toward the central lobe for the high CR group (red circles). In order to quantify the differences between groups, we obtained the average difference of the node centrality (∆𝒆𝒗(𝒊)), within-module degree ∆𝒛(𝒊) and participation coefficient ∆𝒑(𝒊) of each node for both groups. Figure 3 shows the significant differences obtained in node centrality (A), within-module degree (B) and participation coefficient (C) between groups (𝒑𝒗𝒂𝒍 ≤ 0.05). The high CR group showed higher values of the eigenvector centrality in nodes over the central lobe, near the parieto-occipital sulcus, as well as some tiny regions in the frontal lobe (i.e., 𝒆𝒗(𝒊)𝒉𝒊𝒈𝒉 > 𝒆𝒗(𝒊)𝒍𝒐𝒘 and subsequently, ∆𝒆𝒗(𝒊) > 𝟎). Conversely, cortical tissue placed at the left- temporal and occipital lobes have higher centrality in the low CR population (i.e., ∆𝒆𝒗(𝒊) < 0). High CR group presented higher within-module degree over both temporal brain areas and in the occipital lobe, while low CR group showed higher within-module degree in some regions located at the central lobe. Finally, there were few statistical differences between groups in the participation coefficient, showing the high CR group higher values than the low CR one, especially at regions placed at the frontal lobe. Statistical analysis based on a Man-Whitney U test shows similar results (see Supp. Info. for details). It is worth mentioning that both the eigenvector centrality and the within-module degree are quantifying the percentage of importance a node has with regard to the other nodes of the network. Therefore, they should be interpreted as a way of evaluating the increase/decrease of importance with respect to all nodes, in the case of eigenvector centrality, or the nodes belonging to the same lobe, in the case of the within-module degree. Note that, if nodes belonging to the temporal and occipital lobes have higher centrality, it should be at expenses of other nodes of the functional network (in this case, those located at the central lobe), since the sum of the centralities of all nodes is constant (due to the fact that the modulus of the eigenvector is one). In other words, when comparing both groups, it would not be possible to observe higher centrality for all nodes of a given group (High CR or Low CR). Figure 3. Differences between high and low CR groups at the node level. Black dots indicate the Euclidean position of the 148 magnetometers (nodes). Circles filled with lilac colour show nodes with significant statistical differences in: Eigenvector centrality ∆ev(i) (A), within-module degree z-score ∆z(i) (B) and participation coefficient ∆p(i) (C). Green borders correspond to those nodes that have greater values in the high CR group, and red borders represent those nodes that exhibit greater values in the low CR group. Circle sizes are proportional to the absolute value of the differences between groups. Here we present the results obtained with the permutation test. Results obtained with Mann-Whitney U-test are similar (see Fig. S3 of Supp. Info. for details). Macro-scale: Analysing the topology of the whole network We analysed how the average network parameters were modified according to the level of CR of the participants. First, we calculated the mean and standard deviation of six network parameters for each subject (network strength 𝑆, outreach 𝑂, weighted clustering coefficient 𝐶𝑤, average neighbour strength 𝑆𝑛𝑛, global efficiency 𝐸𝑔 and average shortest path 𝑑). Second, we carried out a permutation test and a nonparametric Mann-Whitney U-test (see Section Statistical analysis for details) for each network metric. Finally, those metrics with p-values<0.05 were accepted to have statistically significant differences. In our results, the low CR group exhibited higher averages values than the high CR group in all parameters calculated, except for 𝑑 (see Fig. 4 for details). However, only the network outreach 𝑂 showed statistical significant differences both in the permutation (p= 0.0232) and the rank-sum (p=0.039) tests. Interestingly, the low CR group showed a higher 𝑂 (𝑂 low= 13.57 and 𝑂 high=12.29). It is also remarkable that the average shortest path length 𝑑 exhibit significant differences (p=0.0491) at the rank sum test, despite not passing the permutation test. In this case the average values of shortest path in low CR are lower than in the high CR group (𝑑low= 12.67 and 𝑑high=14.13) (see also Tab. S1 of the Supp. Info. for the values the averages). Low CR High CR Low CR High CR Low CR High CR Figure 4. Violin plots showing the distribution of subject's features and means for the six network parameters analysed: Network strength S, outreach O, weighted clustering coefficient Cw, average neighbour strength Snn, global efficiency Eg and average shortest path d. Red and blue plots correspond to the low and high CR groups, respectively. Red crosses indicate the mean value. Outreach O (B) showed significant differences between groups both with the permutation test as well as with A Mann-Whitney U-test (indicated by two stars). Average shortest path d (F) showed statistical significant differences between groups only with the latter test (one star). Both statistical analyses were performed with pval < 0.05. Dynamical Analysis of MEG Time Series We found a group of nodes showing statistical differences between the high and low CR groups, both for H(i) and C(i) (𝒑𝒗𝒂𝒍 < 0.05). In accordance with previous sections, the occipital lobe is the brain region that concentrates the highest amount of nodes with statistical differences (see Figures 5 and 6). We obtained a group of 23 nodes whose entropy was higher in the low CR group (except for one node) (see Figure 5). On the contrary, we obtained lower complexity of the signals in the low CR group (except for two nodes), but in a very similar brain localization that in the case of entropy (only two nodes did not overlap) (see Figure 6). These results indicate that a higher level of CR could be associated to a brain dynamics with lower entropy and higher complexity. Figure 5. Differences of entropy H(i) between low and high CR groups. In A, we show the mean and standard deviations of the entropy of nodes that have statistical significant differences between groups. H(i)lowCR (red squares) is higher than H(i)highCR (blue circles) for nearly all nodes. In B, we plot the position of nodes with statistical differences. Node sizes are proportional to ∆H(i) = H(i)highCR−H(i)lowCR. Red borders represent ∆H(i) < 0, otherwise node borders are green. In this case, we used a permutation test but no relevant changes appear when a rank- sum test is carried out (see Supp. Info.). Figure 6. Differences of complexity C(i) between the Low and High CR groups. In A, we show the mean and standard deviations of the entropy of nodes that have statistical significant differences between groups. C(i)lowCR (red squares) is lower than C(i)highCR (blue circles) for nearly all nodes. In B, we plot the position of nodes with statistical differences. Node sizes are proportional to ∆C(i) = C(i)highCR−C(i)lowCR. Green borders represent ∆C(i) > 0, otherwise node borders are red. See Supp. Info. for similar results obtained with the rank-sum test. Correlation between Entropy and Complexity Then, we explored the relationship between these two dynamical properties. Figure 7 shows a complexity-entropy scatter plot for all nodes, with the inset plotting only those nodes with statistical differences. In the figure, each point corresponds to the average of each node over the subjects of a given group. We can observe a significant negative correlation between H(i) and C(i) for low (𝒓 = 0.9874; 𝒑𝒗𝒂𝒍 = 0.0010) and high CR groups (𝒓 = 0.8288; 𝒑𝒗𝒂𝒍r = 0.8288; pval = 0.0010), revealing that those nodes with less entropy are, at the same time, those nodes with higher complexity. Figure 7. Complexity-Entropy Diagram. Diagram of subjects' average complexity vs. entropy for all 148 nodes (i.e., each point represents a node). Red squares correspond to low CR group and blue circles to the high CR group. The inset shows only nodes with statistical significant differences using the permutation test. Specifically, these nodes are the channels that have statistical differences when comparing both the mean entropies and complexities of high vs. low CR groups at node level. See Supp. Info. for similar results obtained with the rank-sum test. Correlations between Topology and Dynamics We explored the relationship between the dynamic properties of the nodes (i.e., 𝐻(𝑖), 𝐶(𝑖)) and two of their topological parameters, namely the node strength 𝑆(𝑖) and the weighted clustering coefficient 𝐶𝑤(𝑖) for each CR group. Figure 8 shows all possible combinations. For the low CR group, we found a positive correlation between 𝐻(𝑖) and 𝑆(𝑖) (𝒓𝟐 = 0.5833; 𝒑𝒗𝒂𝒍 < 0.001), which indicates that nodes with higher strength, i.e. the hubs of the network, are in turn those nodes with higher entropy at their dynamics (see Figure 8A). On the contrary, the correlation between the node complexity and its strength was negative (𝒓𝟐 = 0.7124; 𝒑𝒗𝒂𝒍 < 0.001) (see Figure 8B). The same pattern was found between 𝐻(𝑖) and 𝐶𝑤(𝑖) (positive correlation) (𝒓𝟐 = 0.5713; 𝒑𝒗𝒂𝒍 < 0.001), and between 𝐶(𝑖) and 𝐶𝑤(𝑖) (negative correlation) (𝒓𝟐 = 0.7055; 𝒑𝒗𝒂𝒍 < 0.001) (see Figure 8C-D). Correlations found for the high CR group are qualitatively similar to the low CR but the values 0.740.760.780.80.820.840.860.880.90.920.940.10.150.20.250.30.35Low CRHigh CR0.850.90.950.150.20.250.3 of the correlation parameter are systematically lower (See Tab. S2 of Supp. Info. for details). Figure 8. Correlation between the topological and dynamical metrics of the nodes. Correlation plots for the global dynamical properties (entropy H(i) and complexity C(i)) of the MEG time series and their corresponding topological features (strength S(i) and clustering C(i)). Specifically: A) H(i) vs. S(i); B) C(i) vs. S(i); C) H(i) vs. Cw(i) and D) C(i) vs. Cw(i). Red squares correspond to low CR group and blue circles to the high CR group. All statistical tests rejected the null hypothesis when using permutation test as well as with rank-sum test. DISCUSSION In the present study, we explored if CR might play a role in both the topology and the dynamics of the functional brain networks in healthy elders while performing a memory task. To this end, we classified the participants according to a CR index that combined educational level and occupational attainment into two groups: high or low CR level. We firstly focused on the topological characteristics of the two groups, finding differences at microscale (i.e. nodes) and macroscale (i.e. average over all nodes) levels. At the microscale level (i.e., when nodes are analyzed), we observed that the most influential nodes for both groups were mainly placed over the occipital region, although there was a slight displacement towards the central lobe in the group with high CR. Specifically, we found that the low CR group exhibited higher eigenvector centrality over both temporal and occipital areas, while the high CR group presented higher values in central areas. These observations indicate that the importance of regions placed at the occipital and temporal regions is higher at the low CR group. At the same time, since the sum of the eigenvector centrality of all nodes of the network is constant, other nodes unavoidably have lower centrality values to compensate the higher values reported at the occipital and temporal regions. These nodes are placed at the central lobe, where the high CR group has higher centrality when compared with the low CR one. Interestingly, the within-module degree behaves in the opposite way. Comparing Fig. 3 A and B, we can observe that, in general, regions that have higher eigenvector centrality in the high CR group reduce their within-module degree. The explanation of this, in principle, counter-intuitive observation is that nodes have higher eigenvector centrality as a consequence of having higher weights at their long-range connections (i.e., in weight of the functional connections with nodes laying outside of their module). Besides, high CR group showed higher participation coefficient values than the low CR group, especially in regions located at the frontal lobe. These results pointed out that the importance of a node and its connections with their neighbours inside and outside their community (lobe, in our case) may be differently located during the execution of a cognitive task, according to the level of CR. At macroscale level, the low CR group exhibited higher values in different networks parameters compared to the high CR group (network strength, weighted clustering coefficient, average neighbour strength, global efficiency), being this increment statistically significant only in the outreach parameter both for the permutation and the rank-sum tests. Although strength results did not reach statistical significance, the low CR group presented a higher average in this network parameter suggesting that these participants may require higher brain synchronization to perform the memory task as well as the participants with high CR. As a result of this rise of functional connections, the average shortest path showed lower values, however, only the rank-sum test leaded to statistically significant differences between low and high CR groups. Along with the average shortest path, the outreach was the parameter more altered in the low CR group (in this case passing the two statistical tests). The higher outreach observed in the low CR group indicated the existence of a higher amount of long-range connections, implying a greater energetic cost to accomplish the task. Interestingly, this pattern was also described in a study carried out by our group 23, in which MCI and healthy subjects were compared performing the same Sternberg's letter-probe task. Network's results showed that MCI subjects exhibited an increment of strength and outreach values, and a decrement in the average shortest path, as formerly mentioned. Taken together all these findings, and especially the fact that the network outreach (and as a consequence, the energy consumption) is higher in the low CR group we may consider that brain functioning and network organization of those subjects with low CR have some similarities with MCI subjects during the execution of a memory task, and therefore we suggest that CR should be investigated as a possible factor with a remarkable impact on the cognitive status within the healthy aging. However, it is worth mentioning that, in our case, none of the normalized parameters showed significant statistical differences between low and high CR groups, while in MCI subjects showed differences in the normalized parameters indicating that their topology was more random than networks of the control group 23. In our case, since the normalization is made by equivalent random networks, the lack of statistical differences between groups indicates that we can not claim that functional networks of the low CR group have a higher/lower random organization. It is worth mentioning how the synchronized activity along the whole functional networks was negatively correlated with CR during the cognitive task 11. In this sense, individuals with higher CR required less synchronization between brain regions to successfully carry out the memory test. On the contrary, fMRI recordings obtained during resting state reveal negative correlations, i.e., individuals with higher CR maintain higher synchronization in resting state 15,16. This discrepancy suggests different mechanisms behind the creation and organization of the functional networks during resting state and cognitive tasks. Future studies should investigate this phenomenon with resting and task data with MEG. The interpretation of the brain network profile described in the low CR group may be explained as the result of a neural compensatory effect 4. In this case, as both groups did not differ in their memory task accuracy, we may consider it as a successful neural compensation since the low CR group had to engage additional resources to maintain or improve their performance 12. This effect may be related to the Compensation-Related Utilization of Neural Circuits Hypothesis (CRUNCH) 35, which proposes that there is a loss of efficiency in brain networks of aging, when compared with young adults, since additional brain areas are needed to perform cognitive tasks. In the present study, we used a low demanding memory task, but if we would had modified its difficulty, we would had probably found an unsuccessful compensation attempt in the low CR group, as previously described in MCI subjects 36. The cause of this compensatory effect may be the beginning of a network malfunctioning since individuals with low CR exhibited a lower energetic efficiency to perform the memory task, as previously described in MCI subjects 23. In this line, these individuals may exhibit clinical symptoms of pathological aging (i.e. MCI) earlier or more severe than those subjects with a higher CR. Another noticeable novelty explored in the present work was the dynamical properties of the cortical brain regions while performing the memory task. In line with the microscale results, the main differences between both groups were located over the parieto-occipital lobe. In fact, a lower CR level was related with lower complexity and higher entropy. This means that the brain signals of the low CR group presented a pattern more disorganized than the high CR group within the occipital lobe. In addition, the low CR group exhibited the opposite pattern in the right temporal lobe (i.e. less entropy and more complexity). We then may hypothesize that the recruitment of posterior brain areas was not providing an efficient contribution to adequately perform the task, and therefore the individuals with low CR needed to involve right temporal regions to achieved it. We then verified that these two dynamical properties were inversely correlated, independently of the CR level. Finally, one of the most original approaches performed in this study was the exploration of the relationship between the topology (i.e. node strength and weighted clustering coefficient) and the dynamic characteristics of the nodes (i.e. entropy and complexity). We found that in both CR groups, greater values in both topological measures were related with a higher entropy and a lower complexity. These results demonstrated that the most important nodes or hubs within the functional brain networks were those that exhibited more random dynamics, no matter what the level of CR of the individuals. The results obtained in the current study are limited due to the lack of follow-up of the participants and the small sample size, especially in the case of the high CR group. Notwithstanding, it still demonstrated that CR plays a relevant role in the topological and dynamical network's properties in healthy aging. It should be pointed that one of the advantages of conducting the analysis in the sensor space was that MEG signal was not manipulated, affecting neither the complexity nor the entropy. In addition, these results corroborated that network analysis may be a suitable approach for studying CR, and that MEG may be an appropriate tool to detect functional brain changes in healthy aging. Thus, the profiles described in this work for the low CR group indicate a dual pattern of compensation and network impairment. As a consequence, if one of these subjects suffers brain damage or starts a neurodegenerative disease process, their closeness to random network organization would probably be a risk factor for developing a more severe cognitive impairment, although some compensation mechanisms are still present. Thus, this study highlights the importance of preventive interventions, including cognitive training, which could induce a more protective brain network organization in the case of brain damage in elderly subjects. ACKNOWLEDGMENTS This study was supported by the Spanish Ministry of Economy and Competitiveness under Project FIS2013-41057-P, a postdoctoral fellowship from the Complutense University of Madrid to María Eugenia López and a predoctoral fellowship to David López Sanz (BES-2013-063772), both from the Spanish Ministry of Economy and Competitiveness. AUTHOR CONTRIBUTIONS JHM contributed writing paper, in-silico experimental setup, wrote code for results, analysed the data; MEL with MEG recordings and writing paper; PA analysed the data; MC contributed with results and statistical analysis; JAP-P worked on the statistical analysis; DL-S contributed writing the paper; PG worked on the clinical assessment of the participants, FM helped in the design the paradigm and JMB conceived the analysis, analyzed the data and helped writing the manuscript. ADDITIONAL INFORMATION Competing financial interests: The authors declare no competing financial interests. REFERENCES 1. Stern, Y. Cognitive reserve in ageing and Alzheimer's disease. Lancet Neurol. 11, 1006– 12 (2012). 2. Katzman, R. Education and the prevalence of dementia and Alzheimer's disease. Neurology 43, 13–20 (1993). 3. Satz, P. Brain reserve capacity on symptom onset after brain injury: A formulation and review of evidence for threshold theory. Neuropsychology (1993). doi:10.1037/0894- 4105.7.3.273 4. 5. Stern, Y. Cognitive reserve. Neuropsychologia 47, 2015–2028 (2009). Valenzuela, M. J. & Sachdev, P. Brain reserve and dementia: A systematic review. Psychol. Med. 36, 441–454 (2006). 6. Bartrés-Faz, D. & Arenaza-Urquijo, E. M. Structural and functional imaging correlates of cognitive and brain reserve hypotheses in healthy and pathological aging. Brain Topogr. 24, 340–57 (2011). 7. Serra, L. et al. Neuroanatomical correlates of cognitive reserve in Alzheimer disease. Rejuvenation Res. 14, 143–51 (2011). 8. Perneczky, R. et al. Schooling mediates brain reserve in Alzheimer's disease: findings of fluoro-deoxy-glucose-positron emission tomography. J. Neurol. Neurosurg. Psychiatry 77, 1060–1063 (2006). 9. Serra, L. et al. Cognitive reserve and the risk for Alzheimer's disease: a longitudinal study. Neurobiol. Aging 36, 592–600 (2015). 10. López, M. E. et al. Searching for Primary Predictors of Conversion from Mild Cognitive Impairment to Alzheimer's Disease: A Multivariate Follow-Up Study. J. Alzheimers. Dis. 52, 133–143 (2016). 11. Solé-Padullés, C. et al. Brain structure and function related to cognitive reserve variables in normal aging, mild cognitive impairment and Alzheimer's disease. Neurobiol. Aging 30, 1114–24 (2009). 12. López, M. E. et al. Cognitive reserve is associated with the functional organization of the brain in healthy aging: A MEG study. Front. Aging Neurosci. 6, 1–9 (2014). 13. Bullmore, E. & Sporns, O. Complex brain networks: graph theoretical analysis of structural and functional systems. Nat. Rev. Neurosci. 10, 186–98 (2009). 14. Wook Yoo, S. et al. A Network Flow-based Analysis of Cognitive Reserve in Normal Ageing and Alzheimer's Disease. Sci. Rep. 5, 10057 (2015). 15. Marques, P. et al. The functional connectome of cognitive reserve. Hum. Brain Mapp. 37, 3310–3322 (2016). 16. van den Heuvel, M. P., Mandl, R. C. W., Kahn, R. S. & Hulshoff Pol, H. E. Functionally linked resting-state networks reflect the underlying structural connectivity architecture of the human brain. Hum. Brain Mapp. 30, 3127–41 (2009). 17. deToledo-Morrell, L. et al. A 'stress' test for memory dysfunction. Electrophysiologic manifestations of early Alzheimer's disease. Arch. Neurol. 48, 605–9 (1991). 18. Latora, V. & Marchiori, M. Efficient behavior of small-world networks. Phys. Rev. Lett. 87, 198701 (2001). 19. Newman, M. E. J. (Mark E. J. . Networks : an introduction. (Oxford University Press, 2010). 20. Ahnert, S. E., Garlaschelli, D., Fink, T. M. A. & Caldarelli, G. Ensemble approach to the analysis of weighted networks. Phys. Rev. E 76, 16101 (2007). 21. Navas, A. et al. Functional Hubs in Mild Cognitive Impairment. Int. J. Bifurc. Chaos 25, 1550034 (2015). 22. Watts, D. J. & Strogatz, S. H. Collective dynamics of 'small-world' networks. Nature 393, 440–2 (1998). 23. Buldú, J. M. et al. Reorganization of functional networks in mild cognitive impairment. PLoS One 6, e19584 (2011). 24. Oldfield, R. C. The assessment and analysis of handedness: the Edinburgh inventory. Neuropsychologia 9, 97–113 (1971). 25. Lobo, A., Ezquerra, J., Gómez Burgada, F., Sala, J. M. & Seva Díaz, A. [Cognocitive mini-test (a simple practical test to detect intellectual changes in medical patients)]. Actas Luso. Esp. Neurol. Psiquiatr. Cienc. Afines 7, 189–202 (1979). 26. Wechsler, D. Wechsler Memory Scale - Revised. (The Psychological Corporation, 1987). 27. Pfeffer, R. I., Kurosaki, T. T., Harrah, C. H., Chance, J. M. & Filos, S. Measurement of functional activities in older adults in the community. J. Gerontol. 37, 323–9 (1982). 28. Stam, C. J. & van Dijk, B. W. Synchronization likelihood: an unbiased measure of generalized synchronization in multivariate data sets. Phys. D Nonlinear Phenom. 163, 236–251 (2002). 29. Stam, C. J. et al. Magnetoencephalographic evaluation of resting-state functional connectivity in Alzheimer's disease. Neuroimage 32, 1335–44 (2006). 30. Guimerà, R. & Amaral, L. A. N. Cartography of complex networks: modules and universal roles. J. Stat. Mech. Theory Exp. 2005, P02001 (2005). 31. Bandt, C. & Pompe, B. Permutation entropy: a natural complexity measure for time series. Phys. Rev. Lett. 88, 174102 (2002). 32. Rosso, O. A., Larrondo, H. A., Martin, M. T., Plastino, A. & Fuentes, M. A. Distinguishing noise from chaos. Phys. Rev. Lett. 99, 154102 (2007). 33. López-Ruiz, R., Mancini, H. L. & Calbet, X. A statistical measure of complexity. Phys. Lett. A 209, 321–326 (1995). 34. Zanin, M., Zunino, L., Rosso, O. A. & Papo, D. Permutation Entropy and Its Main Biomedical and Econophysics Applications: A Review. Entropy 14, 1553–1577 (2012). 35. Reuter-Lorenz, P. a. & Cappell, K. a. Neurocognitive aging and the compensation hypothesis. Curr. Dir. Psychol. Sci. 17, 177–182 (2008). 36. López, M. E. et al. Synchronization during an internally directed cognitive state\nin healthy aging and mild cognitive impairment: a MEG study. Age (Omaha). 36, 9643 (2014). SUPLEMENTARY INFORMATION S1.- DEFINITION OF NETWORK METRICS The strength s(i) of a node i is the sum of the weights 𝑤𝑖𝑗 of all its links: 𝒔(𝒊) = ∑ 𝒘𝒊𝒋 𝒋∈𝑵 The strength S of a functional network is the average of the strength of all its nodes. The strength of the nearest neighbours snn(i) as the average strength of all neighbours of node i. The outreach o(i) of a node i is the sum of the links' weights 𝑤𝑖𝑗 multiplied by the link Euclidean lengths 𝑙𝑖𝑗 between node i and node j: 𝒐(𝒊) = ∑ 𝒍𝒊𝒋𝒘𝒊𝒋 𝒋∈𝑵 The outreach O of a functional network is the average of the outreach of all its nodes. The weighted clustering coefficient cw(i) of a node i quantifies the percentage of neighbours of a certain node that, in turn, are neighbours between them, taking into account the weight of the connections: 𝒄𝒘(𝒊) = ∑ 𝒘𝒊𝒋𝒘𝒋𝒌𝒘𝒊𝒌 𝒋𝒌 ∑ 𝒘𝒊𝒋𝒘𝒊𝒌 𝒋𝒌 The weighted clustering coefficient Cw of a functional network is the average of the weighted clustering coefficient of all its nodes. The eigenvector centrality ev(i) of a node i measures the importance of a node according to the importance of its neighbours. It is obtained from the eigenvector associated with the largest eigenvalue of the connectivity matrix. The within-module degree z(i) of a node i measures the importance of a node inside its community. In our case, the community of a node is the lobe the node belongs to: 𝒛(𝒊) = 𝒌𝒊(𝒎𝒊) − 〈𝒌𝒊(𝒎𝒊)〉 𝝈𝒌(𝒎𝒊) where 𝒌𝒊(𝒎𝒊) is the degree of node i inside its community (lobe), and 〈𝒌𝒊(𝒎𝒊)〉 and 𝝈𝒌(𝒎𝒊) are the average and the standard deviation of the degree inside the community, respectively. The participation coefficient p(i) of node i quantifies the percentage of links of a node that reach other communities: 𝒑(𝒊) = 𝟏 − ∑( )𝟐 𝒌𝒊(𝒎) 𝒌𝒊 𝒎 where 𝒌𝒊(𝒎) is the degree of node i inside community m. The average shortest path d of node i is the average of the minimum number of nodes to be visited when going from node i to j. To obtain d, we weight the distances between nodes Dij, as the inverse of the elements of the connectivity matrix 𝒘𝒊𝒋, i.e. Dij = 1/𝒘𝒊𝒋. Next, we calculate the shortest- path distance between every pair of nodes using the Dijkstra's algorithm (Dijkstra, 1959). This way, we obtain the shortest-path matrix 𝒅𝒊𝒔𝒊𝒋 (Newman, 2010), and finally the average shortest path d is obtained as the average of the distance of each node to the rest of the network: 𝒅 = 𝟏 𝑵(𝑵 − 𝟏) ∑ 𝑑𝑖𝑠𝑖𝑗 𝑖≠𝑗 The global Efficiency Eg of a network, first introduced by Latora & Marchiori (2001), overcomes the fact that certain nodes of a network could be isolated from the others, thus leading to infinite distance between them. Mathematically, Eg is defined as the harmonic mean of the inverse of the shortest paths between all nodes of the network, with 𝑑𝑖𝑠𝑖𝑗, being the shortest path between nodes i and j: 𝑬𝒈 = 𝟏 𝑵(𝑵 − 𝟏) 1 𝑑𝑖𝑠𝑖𝑗 ∑ 𝑖≠𝑗 S2.- ENTROPY AND COMPLEXITY In the matter of dynamical complexity, information plays an important role as a feature that describes the outermost bounds of periodicity, chaos and complexity. In this sense, the Bandt & Pompe (2002) BP method obtains the intrinsic temporal symbol sequences {St} from the neighbouring steps of a time series (see Fig. S1 for a qualitative explanation). This symbol sequences depend on an embedding dimension D = 3, 4, 5, …, which represents the amount of past information, being D the number of neighbouring samples. In this way, D characterizes each {Xt} time series along t = 1, 2, …, M samples. To do that, {Xt} is partitioned into (M - D) overlapping vectors of dimension D. Figure S1: Overlapping vectors for the case D = 3. D! is the number of patterns. This way 3! different types 𝜋 of accessible states are presented. The probability of appearance of the ordinal patterns is contained in P. If an ordinal pattern never appears, it is called a forbidden pattern. The greater the dimension D, the more information about the past state of our system, and the longer the vectors are (containing the ordering of a set of D samples). Each of the vectors is assigned to a time t, sliding the vector at every time step, to get a total of (M – D + 1) overlapping vectors. Hence, for each (M - D) vector, the position of the lowest value will be assigned the ordinal value zero. The position of the highest value will correspond to (D - 1) ordinal value (the highest in the ranking). Thus, the following positions in-between the assigned zero and (D - 1) will be assigned by rating the positions of the remaining samples in the respective ordinal values. When all (M – D + 1) different order types in {St} are calculated, it is possible to obtain the probability distribution function (PDF) P(𝜋), quantifying the probability of finding a certain order pattern associated to {Xt}: 𝑷(𝜋) = #{𝒕𝒕 ≤ 𝑴 − 𝑫, (𝒙𝒕+𝟏, … , 𝒙𝒕+𝑫)𝒉𝒂𝒔 𝒂 𝒕𝒚𝒑𝒆 𝜋} 𝑴 − 𝑫 + 𝟏 In previous equation, 𝜋 is a possible ordinal pattern presented in the sequence {St} and # is its number of appearances. Note that each ordinal pattern is a permutation of 𝜋 = (0,1,2, …, D - 1). In other words, D! represents all possible permutations 𝜋 of order D of the number of accessible states (M - D). As an example, consider the case of D = 3 in Fig. 1.8. The number of patterns or accessible states will be D! = 3! = 6 and the possible patterns 𝜋 will be: {(012), (021), (102), (201), (120), (210)}. From the former ones, vectors that appear in {St} are called ordinal patterns of {Xt}, those that do not appear in {St}, but belong to the possible accessible states are called forbidden patterns. Finally, the discrete PDF of the ordinal patterns P =𝑝𝑗, ∀ j = 1,2, …, N ∧ N = D! is calculated. This PDF, obtained from BP method, carries the temporal information of {Xt} by comparing consecutive samples. In other words, this symbolic technique incorporates the causality effects of a short-memory (of D steps) a time series has. Next, we use the PDF of the ordinal patterns to define the Normalized Permutation Entropy H (Band & Pompe, 2002): Definition S1 - Normalized Permutation Entropy H[P]. It is given by the ratio between the entropy S[P] of the ordinal patterns and Smax = S[Pe], being Pe the uniform probability distribution: H[P] = S[P]/Smax Note that, the normalized permutation entropy H[P] is bounded between [0, 1]. Regarding the finite size effects, the normalized permutation entropy H allows to include a uniform distribution Pe = {1/N, 1/N, …, 1/N} making H to be an intensive property. This uniform distribution Pe, also maximizes the associate-system information entropy S[P], i.e., Smax =log(N)= log(D!). This way, the amount of disorder H[P] based on the information measure S[P] associated to P is defined as the Permutation Entropy because it runs over all D! permutations 𝜋 of order D. On the other hand, the insertion of "a priori" equilibrium distribution {Pe} as a correction for the associated entropy, leads to a discrimination between two populations. In other words, we need to evaluate the distance between both distributions P and Pe. This fact makes S[P] not being enough to effectively characterize {Xt} because there could be some ordinal patterns that belong to P as well as to Pe. This distance accounts for the "order" of the system when one of the few ordinal patterns emerges as the preferred one. The disequilibrium between statistical populations will be the measure to distinguish this non-Euclidean distance. We quantify the disequilibrium Q by adopting some statistical distance D between the possible and accessible states of the systems in P and the equilibrium distribution Pe: Definition S2 - Disequilibrium Q[P]. It evaluates the distance between P and Pe as: Q[P] = Q0 D [P, Pe] where Q0 is a normalization constant leading to 0 ≤ Q[P] ≤ 1. A Q ≠ 0 indicates the existence of preferred states among the accessible ones. In this way, the disequilibrium Q[P], discriminates ordinal patterns in P from the uniform distribution Pe. The zero limit or the minimum disequilibrium, implies that the lowest separation of both populations does not distinguish between ordinal patterns coming from both populations. Meanwhile the upper limit, with a high disequilibrium, is related to the fact of the existence of some privileged ordinal patterns in P. Hitherto, both H and Q give some sense of the understanding of what the dynamical properties of the system are. Nevertheless, we are concerned on evaluating the interplay between the order and disorder of a system. Therefore, it is also desirable to complement these measures with some metric quantifying the complexity of the system. In this way, Bandt and Pompe define the statistical complexity of a system as: Definition S3 - Statistical Complexity C[P]. This complexity mesures is defined as the product between the permutation entropy H and the disequilibrium Q: C[P] = H[P] *Q[P] With this definition, the statistical complexity C accomplishes the first requirement since H and Q are intensive quantities. The second requirement is also achieved since, by means of H and Q, we are measuring the disorder of a system and its distance from the equilibrium. Note that the statistical complexity vanishes either if the system is at equilibrium (maximum disorder) or if it is completely ordered (maximal distance from the equilibrium). Figure 1.9 shows a qualitative plot indicating the interplay between the three measures. Figure S2: Statistical measures based on ordinal patterns. Schematic representation of the interplay between the normalized permutation entropy H[P], disequilibrium Q[P] and the statistical complexity C[P], in terms of a system that range from complete order to complete disorder. The previous definitions of H, Q and C are usually known as Generalized Statistical Complexity Measures (SCM). SCM capture either, the essential details of the dynamics that allow discerning among different degrees of periodicity and randomness, as well as all possible degrees of stochasticity when the information of {Xt} is extracted via the BP method. SCM, not only compute randomness, but a wide range of correlation structures, not already offered by a simple entropy analysis. S3.- COMPLEMENTARY RESULTS Figure S3. Differences between high and low CR groups at the node level. Black dots indicate the Euclidean position of the 148 magnetometers (nodes). Circles filled with lilac colour show nodes with significant statistical differences obtained with Mann-Whitney U-test in the: Eigenvector centrality ∆ev(i) (A), within-module degree z-score ∆z(i) (B) and participation coefficient ∆p(i) (C). Green borders correspond to those nodes that have greater values in the high CR group, and red borders represent those nodes that exhibit greater values in the low CR group. Circle sizes are proportional to the absolute value of the differences between groups. Here we present the results obtained with the permutation test. Group 𝑆 𝑂 (**) 𝐶𝑤 𝑆𝑛𝑛 𝐸𝑔 𝑑 (*) Low 13.78 13.57 0.102 13.96 0.096 12.69 High 12.75 12.29 0.093 12.76 0.089 14.12 Table S1. Average network parameters of the low and high CR groups. Specifically, the network strength 𝑆, the outreach 𝑂, the weighted clustering coefficient 𝐶𝑤, the average neighbour strength 𝑆𝑛𝑛, the global efficiency 𝐸𝑔 and the average shortest path 𝑑. Asterisks indicate those metrics with statistically significant differences between groups: one asterisk for the parameters passing the rank-sum test and two asterisks for those parameters that also passed the permutation test. Figure S4. Differences of entropy H(i) between low and high CR groups. In A, we show the mean and standard deviations of the entropy of nodes that have statistical significant differences between groups. H(i)lowCR (red squares) is higher than H(i)highCR (blue circles) for nearly all nodes. In B, we plot the ∆H(i) = position of nodes with statistical differences. Node sizes are proportional H(i)highCR−H(i)lowCR. Red borders represent ∆H(i) < 0, otherwise node borders are green. In this case, significant statistical differences were obtained with Mann-Whitney U-test. to Figure S5. Differences of complexity C(i) between the Low and High CR groups. In A, we show the mean and standard deviations of the entropy of nodes that have statistical significant differences between groups. C(i)lowCR (red squares) is lower than C(i)highCR (blue circles) for nearly all nodes. In B, we plot the position of nodes with statistical differences. Node sizes are proportional to ∆C(i) = C(i)highCR−C(i)lowCR. Green borders represent ∆C(i) > 0, otherwise node borders are red. Statistical significant differences were obtained with Mann-Whitney U-test. Figure S6. Complexity-Entropy Diagram. Diagram of subjects' average complexity vs. entropy for all 148 nodes (i.e., each point represents a node). Red squares correspond to low CR group and blue circles to the high CR group. The inset shows only nodes with statistical significant differences using the permutation test. Specifically, these nodes are the channels that have statistical differences when comparing both the mean entropies and complexities of high vs. low CR groups at node level. Statistical significant differences were obtained with Mann-Whitney U-test. 𝑟2 (𝑆, 𝐻)𝐿𝑜𝑤 (𝑆, 𝐻)𝐻𝑖𝑔ℎ (𝑆, 𝐶)𝐿𝑜𝑤 (𝑆, 𝐶)𝐻𝑖𝑔ℎ (𝐶𝑤, 𝐻)𝐿𝑜𝑤 (𝐶𝑤, 𝐻)𝐻𝑖𝑔ℎ (𝐶𝑤, 𝐶)𝐿𝑜𝑤 (𝐶𝑤, 𝐶)𝐻𝑖𝑔ℎ Lin. 0.5833 0.4105 0.6486 0.4902 0.5719 0.3423 0.6417 0.4246 Pol. 0.5833 0.4075 0.7124 0.5159 0.5713 0.3332 0.7055 0.4733 Table S2. Coefficient 𝑟2 for the correlations of Fig. 8 of the main text. First (second) row represent the linear (second order polynomic) fit. Columns represent the 𝑟2 of both averaged variables (structural and dynamical) for the low and high CR groups. Columns 1 and 2 are for Fig. 8A, columns 3 and 4 are for Fig. 8B, columns 5 and 6 are for Fig. 8C. The last two columns are for Fig. 8D. Permutation test and rank-sum test rejected Ho with p<0.001. REFERENCES 0.740.760.780.80.820.840.860.880.90.920.940.10.150.20.250.30.35Mann-Whitney U-testLow CRHigh CR0.850.90.950.150.20.250.3 Bandt, C. & Pompe, B. Permutation Entropy: A Natural Complexity Measure for Time Series. Phys. Rev. Lett. 88, 174102 (2002). Dijkstra, E.W. A note on two problems in connexion with graphs. Numerische Mathematik 1, 269–271 (1959). Latora, V. & Marchiori, M. Efficient behavior of small-world networks. Physical Review Letters 87, 198701 (2001). Newman, M. E. J. Networks: An introduction. (Oxford University Press, 2010). Garibotto,V., Borroni,B., Sorbi,S., Cappa,S.F., Padovani,A., and Perani,D.(2011). Education and occupation provide reserve in both ApoE ε4 carrier and non carrier patients with probable Alzheimer's disease. Neurol.Sci. 33, 1037–1042. doi:10.1007/s10072-011-0889-5
1209.0426
2
1209
2012-11-21T14:30:49
A biophysical observation model for field potentials of networks of leaky integrate-and-fire neurons
[ "q-bio.NC" ]
We present a biophysical approach for the coupling of neural network activity as resulting from proper dipole currents of cortical pyramidal neurons to the electric field in extracellular fluid. Starting from a reduced threecompartment model of a single pyramidal neuron, we derive an observation model for dendritic dipole currents in extracellular space and thereby for the dendritic field potential that contributes to the local field potential of a neural population. This work aligns and satisfies the widespread dipole assumption that is motivated by the "open-field" configuration of the dendritic field potential around cortical pyramidal cells. Our reduced three-compartment scheme allows to derive networks of leaky integrate-and-fire models, which facilitates comparison with existing neural network and observation models. In particular, by means of numerical simulations we compare our approach with an ad hoc model by Mazzoni et al. [Mazzoni, A., S. Panzeri, N. K. Logothetis, and N. Brunel (2008). Encoding of naturalistic stimuli by local field potential spectra in networks of excitatory and inhibitory neurons. PLoS Computational Biology 4 (12), e1000239], and conclude that our biophysically motivated approach yields substantial improvement.
q-bio.NC
q-bio
A biophysical observation model for field potentials of networks of leaky integrate-and-fire neurons Peter beim Graben∗ Bernstein Center for Computational Neuroscience Berlin, Germany Department of German Language and Linguistics, Humboldt-Universitat zu Berlin Serafim Rodrigues Centre for Robotics and Neural Systems, School of Computing and Mathematics, University of Plymouth, United Kingdom Abstract We present a biophysical approach for the coupling of neural network ac- tivity as resulting from proper dipole currents of cortical pyramidal neurons to the electric field in extracellular fluid. Starting from a reduced three- compartment model of a single pyramidal neuron, we derive an observation model for dendritic dipole currents in extracellular space and thereby for the dendritic field potential that contributes to the local field potential of a neu- ral population. This work aligns and satisfies the widespread dipole assump- tion that is motivated by the "open-field" configuration of the dendritic field potential around cortical pyramidal cells. Our reduced three-compartment scheme allows to derive networks of leaky integrate-and-fire models, which facilitates comparison with existing neural network and observation models. In particular, by means of numerical simulations we compare our approach ∗Department of German Language and Linguistics Humboldt-Universitat zu Berlin Unter den Linden 6 D -- 10099 Berlin Phone: +49-30-2093-9632 Fax: +49-30-2093-9729 Email address: [email protected] (Peter beim Graben) URL: www.beimgraben.info (Peter beim Graben) Preprint submitted to Frontiers in Computational Neuroscience October 15, 2018 with an ad hoc model by Mazzoni et al. [Mazzoni, A., S. Panzeri, N. K. Logothetis, and N. Brunel (2008). Encoding of naturalistic stimuli by local field potential spectra in networks of excitatory and inhibitory neurons. PLoS Computational Biology 4 (12), e1000239], and conclude that our biophysically motivated approach yields substantial improvement. Keywords: biophysics, neural networks, leaky integrate-and-fire neuron, current dipoles, extracellular medium, field potentials 1. Introduction Since Hans Berger's 1924 discovery of the human electroencephalogram (EEG) (Berger 1929), neuroscientists achieved much progress in clarifying its neural generators (Creutzfeldt et al. 1966a,b, Nunez and Srinivasan 2006, Schomer and Lopes da Silva 2011). These are the cortical pyramidal neu- rons, as sketched in Fig. 1, that possess a long dendritic trunk separat- ing mainly excitatory synapses at the apical dendritic tree from mainly in- hibitory synapses at the soma and at the perisomatic basal dendritic tree (Creutzfeldt et al. 1966a, Spruston 2008). In addition, they exhibit an ax- ial symmetry and are aligned in parallel to each other, perpendicular to the cortex' surface, thus forming a palisade of cell bodies and dendritic trunks. When both kinds of synapses are simultaneously active, inhibitory synapses generate current sources and excitatory synapses current sinks in extracellu- lar space, hence causing the pyramidal cell to behave as a microscopic dipole surrounded by its characteristic electrical field, the dendritic field potential (DFP). The densely packed pyramidal cells form then a dipole layer whose superimposed currents give rise to the local field potential (LFP) of neural masses and eventually to the EEG (Lind´en et al. 2010, Lind´en et al. 2011, Nunez and Srinivasan 2006, Schomer and Lopes da Silva 2011). 2 Figure 1: Sketch of a cortical pyramidal neuron with extracellular current dipole between spatially separated excitatory (open bullet) and inhibitory synapses (filled bullet). Neural in- and outputs are indicated by the jagged arrows. Dendritic current I D causes dendritic field potential (DFP). Despite of the progress from experimental neuroscience, theoretically un- derstanding the coupling of complex neural network dynamics to the electro- magnetic field in the extracellular space poses challenging problems; some of them have been addressed to some extent by B´edard et al. (2004), B´edard and Destexhe (2009), and B´edard and Destexhe (2012). In computer simulation studies, neural mass potentials, such as LFP and EEG are most realistically simulated by means of multicompartmen- tal models (Lind´en et al. 2010, Lind´en et al. 2011, Protopapas et al. 1998, Sargsyan et al. 2001). Lind´en et al. (2010) calculated the current dipole mo- mentum of the DFP for single pyramidal and stellate cells, based on several hundreds compartments of the dendritic trees. Their results were in com- pliance with the standard dipole approximation of the electrostatic multi- pole expansion in the far-field (more than 1 mm remote from the dendritic trunk), but they found rather poor agreement with that approximation in the vicinity of the cell body. For comparison they also computed a "two- monopole" model of one synaptic current and its counterpart, the somatic return current, estimated from the current dipole momentum of the whole 3 dendritic tree. This "two-monopole" model, which corresponds to an electri- cally equivalent single dipole model, obtained from the decomposition of the dendrite into two compartments, better approximates the true current dipole momentum in the vicinity of the pyramidal neuron. By superimposing the DFPs of pyramidal cells to the ensemble LFP, Lind´en et al. (2011) found that LFP properties cannot be attributed to the far-field dipole approximation. However, realistic multicompartmental models are computationally too expensive for large-scale neural network simulations. Therefore, various techniques have been proposed and employed to overcome computational complexity. These include networks of point models (i.e. devoid from any spatial representation), based on conductance models (Hodgkin and Huxley 1952, Mazzoni et al. 2008), population density models (Omurtag et al. 2000), or firing rate models (Wilson and Cowan 1972), which can be seen as a sub class of population density models, with uniform density distribution (Chizhov et al. 2007). In these kinds of models, mass potentials such as LFP or EEG are conventionally described as averaged membrane potential. A different class of models are neural mass models (David and Friston 2003, Jansen and Rit 1995, Wendling et al. 2000, Rodrigues et al. 2010), where mass potentials are estimated either through sums (or actually differences) of postsynaptic potentials (David and Friston 2003) or of postsynaptic currents (Mazzoni et al. 2008). In particular, the model of Mazzoni et al. (2008) which is based on Brunel and Wang (2003), recently led to a series of follow-up studies (Mazzoni et al. 2010, 2011) addressing the correlations between numerically simulated and experimen- tally measured LFP/EEG with spike rates by means of statistical modeling and information theoretic measures. In all of the above point models and their extension to population models, it is assumed that the extracellular space is iso-potential and the majority of studies thereby neglect the effect of extracellular resistance. That is, the extracellular space constitutes a dif- ferent and isolated domain with no effect on neuronal dynamics. In this article we extend the ad hoc model of Mazzoni et al. (2008) to- wards a biophysically better justified approach, taking the dipole character of extracellular currents and fields into account. Basically, our model corre- sponds to the "two-monopole", or, equivalent dipole model of Lind´en et al. (2010) which gave a good fit of the DFP close to the cell body of a cor- tical pyramidal neuron. However, we aim to keep the simplicity of the Mazzoni et al. (2008) model in terms of computational complexity, by en- dowing the extracellular space with resistance and by keeping point-like neu- 4 ronal circuits. That is, in our case we do not quite consider point neurons, nor spatially extended models with detailed compartmental morphology, yet an intermediate level of description is achieved. To this end we propose a reduced three-compartmental model of a single pyramidal neuron (Destexhe 2001, beim Graben 2008, Wang et al. 2004), and derive an observation model for the dendritic dipole currents in the extracellular space and thereby for the DFP that contributes to the LFP of a neural population. Interestingly, our reduced three-compartmental model enables us to derive a leaky integrate- and-fire mechanism (as for a point model (Mazzoni et al. 2008)), with addi- tional observation equations for the DFP, which all together allows to study the relationship between spike rates and LFP. Our derivations also nicely map realistic electrotonic parameters to phenomenological parameters considered in Mazzoni et al. (2008). 2. Material and Methods Mazzoni et al. (2008) consider three populations of neurons, namely exci- tatory cortical pyramidal cells (population 1), inhibitory cortical interneurons (population 2) and excitatory thalamic relay neurons (population 3), pass- ing sensory input to the cortex that is simulated by a random (Erdos-R´enyi) graph of K = 4000 pyramidal and L = 1000 interneurons with connection probability P = 0.2. 2.1. Theory We describe the ith cortical pyramidal neuron [Fig. 1] from population 1 via the electronic equivalent (reduced) three-compartment model Fig. 2 (Destexhe 2001, beim Graben 2008, Wang et al. 2004), which is parsimonious to derive our observation model: one compartment for the apical dendritic tree, another one for soma and perisomatic basal dendritic tree (Lind´en et al. 2010), and the third -- actually a leaky integrate-and-fire (LIF) unit -- for the axon hillock where membrane potential is converted into spike trains by means of an integrate-and-fire mechanism. 5 apical dendritic tree perisomatic dendritic tree axon hillock exterior Vi ø(cid:129)(cid:129)(cid:129)(cid:129)(cid:129)(cid:129) €(cid:129)(cid:129)(cid:129)(cid:129)(cid:129)(cid:129)(cid:129) ‚ €(cid:129)(cid:129)(cid:129)(cid:129)(cid:129)(cid:129)(cid:129) ‚  €(cid:129)(cid:129)(cid:129)(cid:129)(cid:129)(cid:129) ‚ €(cid:129)(cid:129)(cid:129)(cid:129)(cid:129)(cid:129) ‚ RD i RC i €(cid:129)(cid:129)(cid:129)(cid:129) ‚  €(cid:129)(cid:129)(cid:129)(cid:129) ‚ €(cid:129)(cid:129)(cid:129)(cid:129) ‚  €(cid:129)(cid:129)(cid:129)(cid:129) ‚ €(cid:129)(cid:129)(cid:129)(cid:129)(cid:129) ‚  €(cid:129)(cid:129)(cid:129)(cid:129)(cid:129) ‚ E E i1 E E i2 E E i3 E I i1 E I i2 E I i3 ñ ñ ñ ñ ñ ñ EM ó ó ó ó ó ó (cid:19) RE i1 RE i2 RE i3 RI i1 RI i2 RI i3 Ci (cid:129) RM û Ui ú (cid:13)        €(cid:129)(cid:129)(cid:129)(cid:129) ‚  €(cid:129)(cid:129)(cid:129)(cid:129) ‚ €(cid:129)(cid:129)(cid:129)(cid:129) ‚  €(cid:129)(cid:129)(cid:129)(cid:129) ‚ €(cid:129)(cid:129)(cid:129)(cid:129)(cid:129) ‚  €(cid:129)(cid:129)(cid:129)(cid:129)(cid:129) ‚ I B i û €(cid:129)(cid:129)(cid:129)(cid:129)(cid:129)(cid:129)(cid:129) ‚ €(cid:129)(cid:129)(cid:129)(cid:129)(cid:129)(cid:129)(cid:129) ‚€(cid:129)(cid:129)(cid:129)(cid:129)(cid:129)(cid:129) ‚ ú ú  €(cid:129)(cid:129)(cid:129)(cid:129)(cid:129)(cid:129) ‚ interior RA i I D i I HH i RB i Figure 2: Proposed electronic equivalent circuit for a pyramidal neuron (reduced three compartmental model). Note that the apical and basal dendrites are not true compart- ments since capacitors are not explicitly represented, rather, these are implicitly taken into account via EPSP and IPSP static functions, thus keeping computational complexity low. Excitatory synapses are represented by the left-most branch, where exci- tatory postsynaptic potentials (EPSP) at a synapse between a neuron j from population 1 or 3 and neuron i act as electromotoric forces EE ij. These poten- tials drive excitatory postsynaptic currents (EPSC) I E ij, essentially consisting of sodium ions, through the cell plasma with resistance RE ij from the synapse towards the axon hillock. The middle branch describes the inhibitory synapses between a neuron k from population 2 and neuron i. Here, inhibitory postsynaptic potentials (IPSP) EI ik provide a shortcut between the excitatory branch and the trigger zone, where inhibitory postsynaptic currents (IPSC) I I ik (essentially chloride ions) close the loop between the apical and perisomatic dendritic trees. The resistivity of the current paths along the cell plasma is given by RI ik. The cell membrane at the axon hillock itself is represented by the branch at the right hand side. Here, a capacitor Ci reflects the temporary storage capacity of the membrane. The serial circuit consisting of a battery EM 6 „ „ „ „ ƒ … „ ƒ … ƒ … „ ƒ … „ „ „ „ ƒ … „ ƒ … ƒ … „ ƒ … „ ƒ … ƒ … „ ƒ … „ „ „ „ ƒ … „ ƒ … ƒ … „ ƒ … „ „ „ „ ƒ … „ „ „ „ „ ƒ … „ „ „ „ ƒ … „ „ „ „ ƒ … „ „ „ ƒ … „ „ „ „ ƒ … „ ƒ … ƒ … „ ƒ … „ „ „ „ ƒ … „ ƒ … ƒ … „ ƒ … „ ƒ … ƒ … „ ƒ … „ „ „ „ ƒ … „ „ „ „ „ „ „ „ „ and a resistor RM denotes the Nernst resting potential and the leakage con- ductance of the membrane, respectively (Johnston and Wu 1997). Finally, a spike generator (Hodgkin and Huxley 1952, Mazzoni et al. 2008) (indicated by a "black box") is regarded of having infinite input impedance. Both, EPSP and IPSP result from the interaction of postsynaptic receptor kinetics with dendritic low-pass filtering in compartments one and two, respectively (Destexhe et al. 1998, Lind´en et al. 2010). Hence the required capacitances, omitted in Fig. 2, are already taken into account by EE ik. Therefore, we refer to our model as to a "reduced compartment model" here. ij, EI The three compartments are coupled through longitudinal resistors, RA where RA extracellular space (Holt and Koch 1999). i denote the resistivity of the cell plasma and RC i , RB i , RD i , RB i that of i , RC i , RD i Finally, the membrane voltage at the axon hillock Ui (the dynamical state variable) and the DFP Vi, which measures the drop in electrical potential along the extracellular resistor RD i are indicated. For the aim of calculation, the mesh currents I D i (the basal current) and I IF i (the integrate-and-fire current) are indicated. i (the dendritic current), I B The circuit in Fig. 2 obeys the following equations: p I D i = I B i = q I E ij I I ik Xj=1 Xk=1 i − I B dUi + dt ijI E ij + (RA ikI I ik + (RB i I D i . Ui − EM RM i + RD i + RC i = I D I IF i I IF i = Ci EE ij = RE EI ik = RI Vi = RD i )I D i )I IF i + RC i + (RB i )I IF i + Ui , 1 ≤ k ≤ q i + Ui , 1 ≤ j ≤ p (1) (2) (3) (4) (5) (6) (7) Here, p is the number of excitatory and q is the number of inhibitory synapses connected to neuron i. The circuit described by Eqs. (1 -- 7) shows that the neuron i is likely to fire when the excitatory synapses are activated. Then, the integrate-and-fire current I IF i . If, by contrast, also the inhibitory synapses are active, the dendritic current I D is shunted between the apical i i equals the dendritic current I D 7 and perisomatic basal dendritic trees and only a portion could evoke spikes at the trigger zone [Eq. (4)]. On the other hand, the large dendritic current I D i flowing through the extracellular space of resistance RD i , gives rise to a large DFP Vi. In order to simplify the following derivations, we gauge the resting po- tential [Eq. (4)] to EM = 0, yielding Ui RM . From (5) we obtain the individual EPSC's as I IF i = Ci dUi dt + I E ij = 1 RE ij (cid:2)EE ij − (RA i + RD i )I D i − (RB i + RC i )I IF i − Ui(cid:3) . And accordingly, the individual IPSC's from (6) I I ik = 1 RI ik (cid:2)EI ik − (RB i + RC i )I IF i − Ui(cid:3) . Inserting (9) into (1) yields the excitatory dendritic current I D i = p Xj=1 1 RE ij EE ij − gE i [(RA i + RD i )I D i + (RB i + RC i )I IF i + Ui] , where we have introduced the excitatory dendritic conductivity gE i = p Xj=1 1 RE ij . (8) (9) (10) (11) (12) Likewise we obtain the inhibitory dendritic currents from (2) and (10) as I B i = q Xk=1 1 RI ik ik − gI EI i[(RB i + RC i )I IF i + Ui] , with the inhibitory dendritic conductivity gI i = 1 RI ik . q Xk=1 8 (13) (14) With these results, we obtain an interface equation for an observation model as follows. Rearranging (11) yields I D i [1 + gE i (RA i + RD i )] = p Xj=1 1 RE ij ij − gE EE i [(RB i + RC i )I IF i + Ui] (15) Next, we eliminate I IF i through (8): I D i (cid:2)1 + gE i (cid:0)RA i + RD Division by 1 + gE i + RD cellular dendritic dipole current: RB i + RC i RM (cid:19)(cid:21) . i (cid:1)(cid:3) = i (cid:0)RA p 1 RE ij i + RC ij − gE EE i (cid:20)Ci(cid:0)RB + Ui(cid:18)1 + Xj=1 i (cid:1) gives the desired expression for the extra- dUi dt i (cid:1) I D i = p Xj=1 αijEE ij − βi dUi dt − γiUi , with the following electrotonic parameters αij = βi = γi = i )] 1 i (RA i + RD i + RC i ) i + RD i ) ij[1 + gE RE i (RB CigE 1 + gE i (RA i (RM + RB gE RM[1 + gE i (RA i + RC i ) i + RD i )] . (16) (17) (18) (19) In order to derive the evolution equation we consider the integrate-and- fire current I IF that is given through (3). The individual EPSCs and IPSCs i have already been obtained in (9) and (10), respectively. Inserting (13) into (3) yields I IF i [1 − gI i(RB i + RC i )] − gI iUi = I D i − q Xk=1 1 RI ik EI ik . Next we insert our interface equation (16) and also (8): dUi dt (cid:20)Ci + Ui RM(cid:21) [1 − gI i(RB i + RC i )] − gI iUi = αijEE ij − βi dUi dt − γiUi − p Xj=1 q Xk=1 1 RI ik EI ik 9 and obtain after some rearrangements {Ci(cid:2)1 − gI i(RB i + RC i )(cid:3) + βi} dUi dt + p 1 − gI Xj=1 i(RB i + RM) + RMγi i + RC RM q αijEE ij − Xk=1 1 RI ik EI ik and after multiplication with ri = RM i + RC i + RM) + RMγi 1 − gI i(RB the dynamical law for the membrane potential at axon hillock: τi dUi dt + Ui = p Xj=1 wE ij EE ij − q Xk=1 wI ik EI ik , where we have introduced the following parameters: • time constants τi = ri{Ci(cid:2)1 − gI i(RB i + RC i )(cid:3) + βi} • excitatory synaptic weights • inhibitory synaptic weights wE ij = riαij wI ik = ri RI ik . Ui = (20) (21) (22) (23) Using the result (20), we can also eliminate the temporal derivative in the interface equation (16) through dUi dt = 1 τi " p Xj=1 wE ij EE ij − q Xk=1 wI ik EI ik − Ui# (24) which yields I D i = p Xj=1(cid:18)αij − βi τi wE ij(cid:19) EE ij + βi τi q Xk=1 wI ik EI ik +(cid:18) βi τi − γi(cid:19) Ui . 10 And eventually, by virtue of Eq. (7) after multiplication with RD field potential (DFP) i the dendritic Vi = p Xj=1 with parameters wE ij EE ij + q Xk=1 wI ik EI ik + ξiUi , wE ij = RD i wE ij(cid:18) 1 ri − βi τi(cid:19) wI ik = RD i wI ik βi τi ξi = RD i (cid:18) βi τi − γi(cid:19) . (25) (26) (27) (28) The change in sign of the inhibitory contribution from Eq. (20) to Eq. (25) has an obvious physical interpretation: In (20), the change of membrane potential Ui and therefore the spike rate is enhanced by EPSPs but diminished by IPSPs. On the other hand, the dendritic shunting current I D in (25) is large i for both, large EPSPs and large IPSPs. From Eq. (20) we eventually obtain the neural network's dynamics by taking into account that postsynaptic potentials are obtained from presy- naptic spike trains through temporal convolution with postsynaptic impulse response functions, i.e. EEI ij (t) =Z t −∞ sEI i (t − t′)Rj(t′) dt′ (29) where sEI tions, respectively, and Rj is the spike train i (t) are excitatory and inhibitory synaptic impulse response func- Rj(t) =Xtν δ(t − tν − τL) (30) coming from presynaptic neuron j, when spikes were emitted at times tν. The additional time constant τL is attributed to synaptic transmission delay (Mazzoni et al. 2008). These events are obtained by integrating (20) with initial condition Ui(tν) = E. 11 (31) Where E is some steady-state potential (Mazzoni et al. 2008). t = tν the membrane reaches a threshold If at time Ui(t) ≥ θi(t) (32) (with possibly a time dependent activation threshold θi(t)) from below dUi(t) dt > 0 then an output spike δ(t − tν) is generated, which is then followed by a potential resetting as follows Ui(tν+1) ← E . (33) Additionally, the integration of the dynamical law is restarted at time t = tν+1 + τrp after interrupting the dynamics for a refractory period τrp. Inserting (29) into (20) entails the evolution equation of the neural net- work τi dUi dt + Ui = Xj=1 Xk=1 p q wE ij sE i (t) ∗ Rj(t) + wI ik sI i(t) ∗ Rk(t) , (34) where the signs had been absorbed by the synaptic weights, such that wE for excitatory synapses and wI ik < 0 for inhibitory synapses, respectively. ij > 0 Following Mazzoni et al. (2008) an individual postsynaptic current I EI ij at a synapse between neurons i and j obeys τ EI d τ EI r dI EI ij dt dxEI ij dt + I EI ij = xEI ij + xEI ij = F EI ij , (35) (36) d are decay time constants and τ EI where τ EI and IPSC, respectively. Auxiliary variables are denoted by xEI prescribes presynaptic forcing are rise time constants of EPSC ij , while F EI ij r F EI ij = τiJijRj(t) (37) with spike train (30). Here, Jij = vwEI as voltage unit. ij denotes synaptic gain with v = 1 mV Note that (37) is essentially a weighted sum of delta functions, such that a single spike can be assumed as particular forcing F = F0δ(t) , 12 (38) with some constant F0. Derivating (35) and eliminating xEI ij transforms Eqs. (35 -- 36) into a linear second-order differential equation with constant coefficients τ EI d τ EI r d2I EI ij dt2 + (τ EI d + τ EI r dI EI ij dt ) + I EI ij = F EI ij . (39) Equation (39) with the particular forcing (38) is solved by a Green's i (t) such that the general solution of (39) is obtained as the function sEI temporal convolution I EI ij (t) =Z t −∞ sEI i (t − t′)F EI ij (t) dt′ . (40) For t 6= 0, (39) assumes its homogeneous form and is easily solved by means of the associated characteristic polynomial τ EI d τ EI r λ2 + (τ EI d + τ EI and λ2 = −1/τ EI r r )λ + 1 = 0 (41) , entailing the Green's functions with roots λ1 = −1/τ EI sEI d i (t) =(cid:16)AEIet/τ EI r − BEIet/τ EI d (cid:17) Θ(t) (42) with the Heaviside step function Θ(t). The constants AEI, BEI > 0 are obtained from the initial conditions sEI i (t) = 0, reflecting causality, and a suitable normalization Z ∞ 0 sEI i (t)dt = 1 . The initial condition yields AEI = BEI ≡ SEI, while the remaining constant SEI = 1 , d − τ EI τ EI r due to normalization. Therefore, the normalized Green's functions are those of Brunel and Wang (2003) sEI i (t) = v τi r (cid:16)et/τ τ EI d − τ EI EI r − et/τ EI d (cid:17) Θ(t) . (43) 13 Now, we are able to compare our DFP Vi [Eq. (25)] with the estimate of Mazzoni et al. (2008) which is given (in our notation) as the sums of the moduli of excitatory and inhibitory synaptic currents, i.e. V MPLB i =Xj ij +Xk I E I I ik (44) where "MPLB" refers to the authors Mazzoni et al. (2008). From (25) and (44), respectively, we compute two models of the local field potential (LFP). First, by summing DFP across all pyramidal neurons (beim Graben and Kurths 2008, Mazzoni et al. 2008), and, second by taking the DFP average (Nunez and Srinivasan 2006), which yields L2 = L1 = Xi K Xi L3 = Xi K Xi L4 = Vi 1 1 V MPLB i V MPLB i Vi , (45) (46) (47) (48) where K is number of pyramidal neurons. 2.2. Parameter estimation Next, we relate the electrotonic parameters of our model to the phe- nomenological parameters of Mazzoni et al. (2008). To this end, we first report their synaptic efficacies in Tab. 1. Synaptic efficacies / mV on interneurons GABA recurrent cortical AMPA external thalamic AMPA 2.7 0.7 0.95 on pyramidal neurons 1.7 0.42 0.55 Table 1: Parameters laid as in Mazzoni et al. (2008). From these, we compute the synaptic weights through wE ij = J E ij/v =(cid:26) 0.42 if 0.55 if j "cortical" j "thalamic" (49) 14 and wI ik = J I ik/v = 1.7 Next, we determine the factors ri by virtue of Eq. (23) through ri = wI ik ¯gGABA = 1.7 1 nS = 1.7 GΩ using the inhibitory synaptic conductivity ¯gGABA = 1 nS, Correspondingly, Eq. (22) allows us to express αij in terms of the excitatory synaptic weights through αij = wE ij ri =(cid:26) 0.25 nS if 0.32 nS if j "cortical" j "thalamic" From αij we can determine the total excitatory synaptic conductivities gE i according to Eq. (17) through αij = 1 i (RA i + RD i )] ij[1 + gE RE p αij Xj=1 gE i "1 − (RA i + RD i ) p Xj=1 αij# = Pp 1 − (RA i + RD gE i = j=1 αij i )Pp 1 j=1 αij αij[1 + gE i (RA i + RD i )] and hence RE ij = (50) (51) Inserting next (18) into (21) yields τi = riCi 1 + gE i (RA i + RD i ) + (RB i + RC i (RA i ){gE i − gI i + RD i ) 1 + gE i[1 + gE i (RA i + RD i )]} .(52) Equation (52) could constraint the choice of the membrane capacitance Ci by choosing τi = 20 ms (Mazzoni et al. 2008). In order to also determine the DFP parameters (26) -- (28), we finally compute the ratios βi τi = ri{1 + gE i (RA i + RD gE i (RB i ) + (RB i + RC i + RC i ) i ){gE i − gI i[1 + gE i (RA i + RD . i )]}} 15 The remaining electrotonic parameters RM i are es- timated from cell geometries as follows. The resistance R of a volume con- ductor is proportional to its length ℓ and reciprocally proportional to its cross-section A, i.e. i , and RD i , RA i , RB i , RC R = ρ ℓ A (53) where ρ is the (specific) resistivity of the medium. Table 2 shows the resistiv- ities of the three kinds of interest which then allows to evaluate the various volume conductor resistances according to Eq. (53). medium cell membrane (at axon hillock) cell plasma (cytoplasm) extracellular space ρ/Ωcm 5 · 107 200 333 Table 2: Resistivities of cell membrane, cell plasma and extracellular space. Parameters from Kole and Stuart (2012), Rall (1977), Mainen et al. (1995), Gold et al. (2007). Note that the resistivity of the cell membrane has to be related to the constant membrane thickness (≈ 10 nm). We consider a total dendritic length of 2ℓ = 20 µm and a dendritic radius of a = 7 µm, that are generally subjected to variation. Equally, parameters that were allowed to vary are the length and radius of the axon hillock, yet herein we consider a length of 2ℓ = 20 µm and radius of a = 0.5 µm (Destexhe 2001, Kole and Stuart 2012, Mainen et al. 1995). To evaluate the intracellular (RA, RB) and extracellular (RD, RC) resistances, respectively, according to Eq. (53), we consider a simple implementation where the length ℓ is half of the dendritic length (i.e. basal and apical length are symmetrical, but this can be broken). However, the cross sectional area for the cytoplasm is simply A = πa2. Finally, the area of the axon hillock is simply the surface area of a cylinder. In order to also determine the cross-section of extracellular space between dendritic trunks we make the following approximations. We assume that dendritic trunks are parallel aligned cylinders of radius a and length ℓ that are hexagonally dense packed. Then the centers of three adjacent trunks form an equilateral triangle with side length 2a and hence area 2√3a2. The enclosed space is then given by the difference of the triangle area and the 16 area of three sixth circle sectors, therefore Aspace = 2√3a2 − 3 6 πa2 =(cid:18)2√3 − 1 2 π(cid:19) a2 . Hence, the cross-section of extracellular space surrounding one trunk is A = 6Aspace =(cid:16)12√3 − 3π(cid:17) a2 . (54) 2.3. Simulations Subsequently, we implement an identical network to the one considered by Mazzoni et al. (2008) with Brian Simulator, that is a Python based en- vironment (Goodman and Brette 2009). However, the derivations from the previous section enables the possibility of setting a dipole observable that measures the local dendritic field potential (DFP) on each pyramidal neurons, given by Eq. (25). This allows then to define a mesoscopic LFP observable, which can be equated either as averaged DFP or simply given as the sum of DFP, given by Eqs. (45 -- 48). Primarily, we compare our LFP measure L4, proposed as the average of DFP, with the Mazzoni et al. LFP L1 which is defined as the sum of absolute values of GABA and AMPA currents Eq. (44). Additionally, we also compare all possible measures, namely, mean membrane potential 1 L2, sum of DFP L3, and the average of DFP L4. K Pi Ui, Mazzoni et al. LFP L1, average of Mazzoni et al. DFP For completeness, we briefly summarize the description of the network (we refer the reader to Mazzoni et al. (2008) for details). The network mod- els a cortical tissue with leaky integrate-and-fire neurons, composed of 1000 inhibitory interneurons and 4000 pyramidal neurons, which are described by the evolution equation (34). The threshold crossings given by Eq. (32) is considered static with θi = 18 mV and the reset potential E = 11 mV. The refractory period for excitatory neurons is τrp = 2 ms while for inhibitory neurons it is τrp = 1 ms. The network connectivity is random and sparse with a 0.2 probability of directed connection between any pair of neurons. The evolution of synaptic currents, fast GABA (inhibitory) and AMPA (ex- citatory) are described via the second order evolution equations (35 -- 36), which are activated by incoming presynaptic spikes represented by Eq. (30). The latency of the postsynaptic currents is set to τL = 1 ms and the rise and decay times are given by Tab. 3. 17 Synaptic times GABA AMPA on interneurons AMPA on pyramidal neurons τr / ms 0.25 0.2 0.4 τd / ms 5 1 2 Table 3: Synaptic rise (τr) and decay times (τd). Parameters laid as in Mazzoni et al. (2008). ij Moreover, synaptic efficacies, J EI , for simulation were presented in Tab. 1. Note that relation (49) then allows to determine the synaptic weights. Addi- tionally, all neurons receive external thalamic excitatory inputs, that is, via AMPA-type synapses, which are activated by random Poisson spike trains, with a time varying rate that is identical for all neurons. Specifically, the thalamic inputs are the only source of noise, which attempts to account for both cortical heterogeneity and spontaneous activity. This is achieved by modeling a two level noise, where the first level is an Ornstein-Uhlenbeck process superimposed with a constant signal and the second level is a time varying inhomogeneous Poisson process. Thus, we have the following time varying rate, λ(t), that feeds into inhomogeneous Poisson process: τn dn(t) = −n(t) + σnr 2 dt λ(t) = [c0 + n(t)]+ τn η(t) (55) (56) where η(t) represents Gaussian white noise, c0 represents a constant signal (but equally could be periodic or other), and the operation [·] is the threshold- linear function, [x]+ = x if x > 0, [x]+ = 0 otherwise, which circumvents negative rates. The constant signal c0 can range between 1.2 to 2.6 spikes/ms. The parameters of the Ornstein-Uhlenbeck process are τn = 16 ms and the standard deviation σn = 0.4 spikes/ms. For complete exposition, we note that from an implementation viewpoint (within the Brian simulator), a copy of the postsynaptic impulse response function (29) has to be evaluated to calculate the dendritic field poten- tial (DFP) (25) with weights wEI ij . This implies evaluating the second or- der process (35) -- (36) with a different forcing term. Specifically, starting from I EI ij (t) ≡ wEI and pre-multiplying both sides with wEI ij and subsequently re-arranging we obtain the desired forcing term F EI ij = wEI ij with equation (37) and using relation (49) we finally obtain F EI ij τivRj(t). ij . Note further that by expanding the term F EI i (t) ∗ F EI ij (t) = sEI ij = wEI ij F EI ij /wEI ij EEI ij 18 3. Results Following Mazzoni et al. (2008), the network simulations are run for two seconds with three different noise levels, specifically, receiving a constant signal with three different rates 1.2, 1.6 and 2.4 spikes/ms as depicted in Fig. 3. Note that these input rates do not mean that a single neuron fires at these high rates. Rather, it can be obtained from multiple neurons that jointly fire with slower, yet desynchronized, rates converging at the same postsynaptic cell. The Poisson process ensures that this is well represented. 19 e t a r c i m a l a h t r e b m u n n o r u e n c e s / s e k p s i c e s / s e k p s i ) V ( 1 L P F L ) V m ( 4 L P F L Signal rate = 1.2 spikes/ms Signal rate = 1.6 spikes/ms Signal rate = 2.4 spikes/ms 2.0 1.5 1.0 0.5 A 2.5 2.0 1.5 1.0 0.5 0.0 C B 4.0 3.5 3.0 2.5 2.0 1.5 1750 1800 1850 1900 1950 2000 1750 1800 1850 1900 1950 2000 1.0 1750 1800 1850 1900 1950 2000 200 150 100 50 D 200 150 100 50 E 200 150 100 50 F 0 1800 1840 1880 1920 0 1750 1800 1850 1900 1950 2000 0 1750 1800 1850 1900 1950 2000 3.0 2.0 1.0 0.0 Interneurons G 2.0 1.0 0.0 H 4.0 3.0 2.0 1.0 0.0 I 1750 1800 1850 1900 1950 2000 1750 1800 1850 1900 1950 2000 1750 1800 1850 1900 1950 2000 20 15 10 5 J Pyramidal neurons 16 12 8 4 K 30 20 10 L 0 1750 1800 1850 1900 1950 2000 0 1750 1800 1850 1900 1950 2000 0 1750 1800 1850 1900 1950 2000 350 300 250 200 150 100 50 M 350 300 250 200 150 100 50 N 600 500 400 300 200 100 O 0 1750 1800 1850 1900 1950 2000 0 1750 1800 1850 1900 1950 2000 0 1750 1800 1850 1900 1950 2000 14 12 10 8 6 4 P 16 14 12 10 8 6 4 Q 26 24 22 20 18 16 14 12 10 R 2 1750 1800 1850 1900 1950 2000 2 1750 1800 1850 1900 1950 2000 8 1750 1800 1850 1900 1950 2000 Time (ms) Time (ms) Time (ms) Figure 3: Dynamics of the network and LFP comparisons: The three columns represent different runs of the network for three different rates, 1.2, 1.6 and 2.4 spikes/ms. In each column, all panels show the same 250 ms (extracted from 2 seconds simulations). The top panels (A-C) represent thalamic inputs with the different rates. The second top panels (D-F) corresponds to a raster plot of the activity of 200 pyramidal neurons. Panels (G-I) depict average instantaneous firing rate (computed on a 1 ms bin) of interneurons (blue) and panels (J-L) correspond to average instantaneous firing rate of pyramidal neurons. Panels (M-O) show the Mazzoni et al. LFP L1 from Eq. (45). Finally, panels (P-R) depict our proposed LFP measure L4, which is the average of dendritic field potential (DFP) [Eq. (48)]. 20 The focus is to compare our proposed measure L4, defined as mean of the dendritic field potential (DFP) [Eq. (48)], with the Mazzoni et al. LFP L1 from Eq. (45). In Fig. 3 one sees two main striking differences between the two measures, namely in frequency and in amplitude. Specifically, L1 responds instantaneously to the spiking network activity by means of high frequency oscillations. Moreover, L1 also exhibits a large amplitude. In contrast, our mean dendritic field potential L4 measures comparably to ex- perimental LFP, that is, in the order of millivolts, and although it responds to population activity, it has a relatively smoother response. Actually one can realize that our LFP estimate represents low-pass filtered thalamic input. The physiological relevance of this is not yet clear in our work. However, recent work (Poulet et al. 2012) shows that desynchronized cortical state dur- ing active behavior is driven by a centrally generated increase in thalamic action potential firing (i.e. thalamic firing controls cortical states). Thus, it seems that cortical synchronous activity is suppressed when thalamic input increases, thereby suggesting that cortical desynchronized states to be re- lated to sensory processing. This work further quantifies these observations by applying Fast Fourier Transform (FFT) to cortical EEG and subsequently comparing with thalamic firing rate by means of Pearson correlation coeffi- cient. Unfortunately they do not quantify the amount of thalamic oscillations contained within the cortical EEG. Yet, to keep a comparable comparison between measures, we also compute the average of the Mazzoni et al. DFP L2 [Eq. (48)] and additionally the mean membrane potential (the standard considered in the neuroscientific literature). These are shown in Fig. 4. 21 ) V m ( e n a r b m e m n a e m ) V ( 1 L P F L ) V m ( 2 L P F L ) V ( 3 L P F L ) V m ( 4 L P F L Signal rate = 1.2 spikes/ms Signal rate = 1.6 spikes/ms Signal rate = 2.4 spikes/ms 3 A 13 12 11 10 9 8 7 B 12 10 8 6 4 2 C 13 12 11 10 9 8 7 6 5 1750 1800 1850 1900 1950 2000 6 1750 1800 1850 1900 1950 2000 0 1750 1800 1850 1900 1950 2000 350 300 250 200 150 100 50 D 350 300 250 200 150 100 50 F E 600 500 400 300 200 100 0 1750 1800 1850 1900 1950 2000 0 1750 1800 1850 1900 1950 2000 0 1750 1800 1850 1900 1950 2000 80 70 60 50 40 30 20 10 G 90 80 70 60 50 40 30 20 10 0 1750 1800 1850 1900 1950 2000 0 1750 1800 50 45 40 35 30 25 20 15 10 J 60 50 40 30 20 10 I H 140 120 100 80 60 40 20 1850 1900 1950 2000 1750 1800 1850 1900 1950 2000 100 K L 90 80 70 60 50 40 30 1750 1800 1850 1900 1950 2000 1750 1800 1850 1900 1950 2000 1750 1800 1850 1900 1950 2000 14 12 10 8 6 4 M 16 14 12 10 8 6 4 N 26 24 22 20 18 16 14 12 10 O 2 1750 1800 1850 1900 1950 2000 2 1750 1800 1850 1900 1950 2000 8 1750 1800 1850 1900 1950 2000 Time (ms) Time (ms) Time (ms) Figure 4: Comparison of different LFP measures when the network receives constant signal with three different rates (1.2, 1.6 and 2.4 spikes/ms). Again, only 250 ms is represented (extracted from 2 sec simulation). The first plot (A-C) corresponding to the the different rates shows the most widespread LFP measure used in the literature, namely average membrane potential 1 Ui. The second panel (D-F) shows the Mazzoni et al. LFP L1 from Eq. (45). The third panel (G-I) displays the average of the Mazzoni et al. DFP L2 [Eq. (46)]. Similarly, the fourth panel (J-L) shows the total, L3, [Eq. (47)] and the last panel (M-O) depicts the averaged, L4, [Eq. (48)] LFP measure. Note the different amplitude scales between measures. K Pi 22 Clearly, in terms of time profile, the summed and averaged observables are similar within the same class of LFP measures. However, in all cases the Mazzoni et al. LFP L1 exhibits a significantly larger order of magnitude, which diverges substantially from experimental LFP amplitudes, typically varying between 0.5 to 2 mV (Niedermeyer 2005, Lakatos et al. 2005). In contrast, although the mean DFP is not contained within the interval from 0.5 to 2 mV it arguably performs better. However, we do concede further work is required. Some gains in improving the different LFP measures can be achieved by applying for example a weighted average, which would mimic the distance of an electrode to a particular neuron by means of a lead field kernel (Nunez and Srinivasan 2006). For example a convolution of either L1 or L2 with a Gaussian kernel (representing the distance to a neuron), would yield a measure that captures better the local field potential or better the dendritic field potential of the nearest neurons. However, further work will be required to properly quantify the gain when space is taken into account. In Fig. 5 we finally contrast the power spectra of the different LFP mea- sures. 23 20 0 -20 -40 -60 -80 z H / B d -100 -120 0 30 20 10 0 -10 -20 -30 -40 -50 -60 0 20 0 -20 -40 -60 -80 0 20 0 -20 -40 -60 -80 z H / B d z H / B d z H / B d -100 0 20 0 -20 -40 -60 -80 z H / B d -100 0 Signal rate = 1.2 spikes/ms Signal rate = 1.6 spikes/ms Signal rate = 2.4 spikes/ms B 1000 2000 3000 4000 5000 E 1000 2000 3000 4000 5000 H A mean membrane potential 1000 2000 3000 4000 5000 D LFP L1 1000 2000 3000 4000 5000 G LFP L2 20 0 -20 -40 -60 -80 -100 -120 0 30 20 10 0 -10 -20 -30 -40 -50 -60 0 20 10 0 -10 -20 -30 -40 -50 -60 -70 1000 2000 3000 4000 5000 0 1000 2000 3000 4000 5000 J LFP L3 1000 2000 3000 4000 5000 M LFP L4 1000 2000 3000 4000 5000 20 0 -20 -40 -60 -80 -100 0 20 0 -20 -40 -60 -80 -100 -120 0 K 1000 2000 3000 4000 5000 N 1000 2000 3000 4000 5000 Hz Hz 0 -20 -40 -60 -80 -100 -120 0 40 20 0 -20 -40 -60 0 30 10 -10 -30 -50 -70 0 40 20 0 -20 -40 -60 -80 -100 0 20 0 -20 -40 -60 -80 -100 0 C 1000 2000 3000 4000 5000 F 1000 2000 3000 4000 5000 I 1000 2000 3000 4000 5000 L 1000 2000 3000 4000 5000 O 1000 2000 3000 4000 5000 Hz Figure 5: Comparison of power spectra of the various LFP measures when the network receives constant signal with three different rates (1.2, 1.6 and 2.4 spikes/ms): The first plot (A-C) corresponding to the different rates shows the power spectrum of the average membrane potential 1 Ui. The second plot (D-F) and third plots (G-I) show power spectra of the total and average of L1 and L2 corresponding to Mazzoni et al. (2008), respectively. The fourth plot (J-L) and fifth plots (M-O) display power spectra of the L3 and L4 measures from our model, respectively. Note we show the full spectrum up to 5 kHz only for convenience due to the fine sample rate. K Pi One interesting feature is that the power spectrum of the Mazzoni et al. LFP measures decays much more slowly that the average membrane potential 24 for higher frequencies. This observation is true for both, L1 and L2. In contrast, our LFP measures L3 and L4 fare better, and in particular, L4 decays at an approximately similar rate as the average membrane potential. 4. Discussion In this article we derived a model for cortical dipole fields, such as den- dritic and local field potential (DFP/LFP) from biophysical principles. To that aim we decomposed a cortical pyramidal cell, the putative generator of those potentials, into three compartments: the apical dendritic tree as the place of mainly excitatory (AMPA) synapses, the soma and the peri- somatic dendritic tree as the place of mainly inhibitory (GABA) synapses and the axon hillock as the place of wave-to-spike conversion by means of an integrate-and-fire mechanism. From Kirchhoff's laws governing an electronic equivalent circuit of our model, we were then able to derive the evolution equation for neural network activity Eq. (34) and, in addition, an observa- tion equation Eq. (25) for the dendritic dipole potential contributing to the LFP of a cortical population. In order to compare our approach with another model discussed in the recent literature (Mazzoni et al. 2008, 2010, 2011) we aligned the parameters of our model with the model of Mazzoni et al. (2008) who approximated DFP as the sum of moduli of excitatory and inhibitory synaptic currents Eq. (44). From both approaches, we computed four different LFP estimates: L1, the sum of Mazzoni et al. DFP, L2, the population average of Mazzoni et al. DFP, L3 the sum of our dipole DFP, and L4 the population average of our dipole DFP [Eqs. (45 -- 48)]. Our results indicate two main effects between our dipole LFP measures and those of Mazzoni et al. Firstly, the measures based on Mazzoni et al. (2008) systematically overestimate LFP amplitude by almost one order of magnitude. One reason for that could be attributed to the direct conversion of synaptic current into voltage without taking extracellular conductivity into account, as properly done in our approach. Yet, another, even more crucial reason is disclosed by our equivalent circuit Fig. 2. In our approach there is just one extracellular current I D flowing from the perisomatic to the apical dendritic tree. In the model of Mazzoni et al. (2008), however, two synaptic currents that might be of the same order of magnitude are superimposed to the DFP. Secondly, the measures based on Mazzoni et al. (2008) also sys- tematically overestimate LFP frequencies. This could probably be attributed 25 partly to spurious higher harmonics introduced by computing absolute val- ues. Moreover, taking the power spectrum shows that the Mazzoni et al. (2008) measure decays much more slowly than the average membrane poten- tial, which is at variance with experimental data. However, at the current stage, both models, that of Mazzoni et al. (2008) and our own, agree with respect to the polarity of DFP and LFP. The mea- sures based on Mazzoni et al. (2008) have positive polarity simply due to the moduli. On the other hand, also the direction of current dipoles in our model is constrained by the construction of the equivalent circuit Fig. 2 where cur- rent sources are situated at the perisomatic and current sinks are situated at apical dendritc tree. Taking this polarity as positive also entails positive DFP and LFP that could only change in strength. However, it is well known from brain anatomy that pyramidal cells appear in at least two layers, III and VI, of neocortex. This is reflected in experiments when an electrode tra- verses different layers by LFP polarity reversals, and, of course, by the fact that LFP and EEG oscillate between positive and negative polarity. Adapt- ing our model to this situation could be straightforwardly accomplished in the framework of neural field theory by fully representing space and simulat- ing layered neural fields (Amari 1977, Jirsa and Haken 1996, beim Graben 2008). By contrast such a generalization is impossible at all with the model of Mazzoni et al. (2008) due to the presence of absolute values. On theses grounds we have good indication that our measure is an im- provement to the Mazzoni et al. LFP measures, and, quite importantly, it is biophysically better motivated than the ad hoc model of Mazzoni et al. (2008). However, much considerable effort is still required to underpin all the relevant LFP mechanisms and to better represent experimental LFP/EEG dynamics. Finally, our work provides a new framework where dendritic field poten- tials and the relationship between firing rates and local fields can be explored without the extreme demand on computational complexity involved in multi- compartmental modeling (Lind´en et al. 2010, Lind´en et al. 2011, Protopapas et al. 1998, Sargsyan et al. 2001) by adopting reduced compartment circuits. For example, we envisage to extend our recent work which maps firing rate model (derived from LIF models) to population density models (Chizhov et al. 2007), but now incorporating our observational DFP model. In addition, our framework is analytically amenable and thus can be applied to any lin- ear differential equation, for instance, GIF (Gif-sur-Yvette Integrate Fire) models, which are improvements to the LIF models and compute more ac- 26 curately spike activations (Rudolph-Lilith et al. 2012). Also resonant mem- branes (mediated by Ca2+ and a Ca2+-activated K+ ionic currents) that describe sub-threshold oscillations and which can be easily expressed by lin- ear equations (Mauro et al. 1970) can be incorporated in our derivations. We note however that our framework can be applied to non-linear equations, with Hodgkin and Huxley (1952) type activation, but it will fall short from explicit and analytical observation equations. Acknowledgements We thank Michelle Lilith, Claude B´edard, Alain Destexhe, and Jurgen Kurths for fruitful discussion. In addition, we would like to thank Samantha Adams for providing help with Brian Simulator installations and initial dis- cussions of Brian usage. This research was supported by a DFG Heisenberg grant awarded to PbG (GR 3711/1-1). References Amari, S.-I. (1977). Dynamics of pattern formation in lateral-inhibition type neural fields. Biological Cybernetics 27, 77 -- 87. B´edard, C. and A. Destexhe (2009). Macroscopic models of local field po- tentials and the apparent 1/f noise in brain activity. Biophysical Jour- nal 96 (7), 2589 -- 2603. B´edard, C. and A. Destexhe (2012). Modeling local field potentials and their interaction with the extracellular medium. In R. Brette and A. Destexhe (Eds.), Handbook of Neural Activity Measurement, pp. 136 -- 191. Cam- bridge: Cambridge University Press. B´edard, C., H. Kroger, and A. Destexhe (2004). Modeling extracellular field potentials and the frequency-filtering properties of extracellular space. Biophysical Journal 86 (3), 1829 -- 1842. beim Graben, P. (2008). Foundations of neurophysics. In P. b. Graben, C. Zhou, M. Thiel, and J. Kurths (Eds.), Lectures in Supercomputational Neuroscience: Dynamics in Complex Brain Networks, Springer Complex- ity Series, pp. 3 -- 48. Berlin: Springer. 27 beim Graben, P. and J. Kurths (2008). Simulating global properties of elec- troencephalograms with minimal random neural networks. Neurocomput- ing 71 (4), 999 -- 1007. Berger, H. (1929). Uber das Elektroenkephalogramm des Menschen. Archiv fur Psychiatrie 87, 527 -- 570. Brunel, N. and X.-J. Wang (2003). What determines the frequency of fast network oscillations with irregular neural discharges? I. synaptic dynamics and excitation-inhibition balance. Journal of Neurophysiology 90 (1), 415 -- 430. Chizhov, A. V., R. S., and T. J. R. (2007). A comparative analysis of a firing-rate model and conductance-based neural population model. Physics Letters A 369, 31 -- 36. Creutzfeldt, O. D., S. Watanabe, and H. D. Lux (1966a). Relations be- tween EEG phenomena and potentials of single cortical cells. I. evoked responses after thalamic and epicortical stimulation. Electroencephalogra- phy and Clinical Neurophysiology 20 (1), 1 -- 18. Creutzfeldt, O. D., S. Watanabe, and H. D. Lux (1966b). Relations between EEG phenomena and potentials of single cortical cells. II. spontaneous and convulsoid activity. Electroencephalography and Clinical Neurophysi- ology 20 (1), 19 -- 37. David, O. and K. J. Friston (2003). A neural mass model for MEG/EEG: coupling and neuronal dynamics. NeuroImage 20, 1743 -- 1755. Destexhe, A. (2001). Simplified models of neocortical pyramidal cells pre- serving somatodendritic voltage attenuation. Neurocomputing 38-40, 167 -- 173. Destexhe, A., F. Mainen, and T. J. Sejnowski (1998). Kinetic models of synaptic transmission. See Koch and Segev (1998), pp. 1 -- 25. Gold, C., D. Henze, and C. Koch (2007). Using extracellular action potential recordings to constrain compartmental models. Journal of Computational Neuroscience 23 (1), 39 -- 58. 28 Goodman, D. and R. Brette (2009). The Brian simulator. Frontiers in Neuroscience 3 (2), 192 -- 197. Hodgkin, A. L. and A. F. Huxley (1952). A quantitative description of mem- brane current and its application to conduction and excitation in nerve. Journal of Physiology 117, 500 -- 544. Holt, G. R. and C. Koch (1999). Electrical interactions via the extracellular potential near cell bodies. Journal of Computational Neuroscience 6, 169 -- 184. Jansen, B. H. and V. G. Rit (1995). Electroencephalogram and visual evoked potential generation in a mathematical model of coupled cortical columns. Biological Cybernetics 73, 357 -- 366. Jirsa, V. K. and H. Haken (1996). Field theory of electromagnetic brain activity. Physical Review Letters 77 (5), 960 -- 963. Johnston, D. and S. M.-S. Wu (1997). Foundations of Cellular Neurophysi- ology. Cambridge (MA): MIT Press. Koch, C. and I. Segev (Eds.) (1998). Methods in Neuronal Modelling. From Ions to Networks (2nd ed.). Computational Neuroscience. Cambridge (MA): MIT Press. Kole, M. and G. Stuart (2012). Signal processing in the axon initial segment. Neuron 73 (2), 235 -- 247. Lakatos, P., A. Shah, K. Knuth, I. Ulbert, G. Karmos, and C. Schroeder (2005). An oscillatory hierarchy controlling neuronal excitability and stim- ulus processing in the auditory cortex. Journal of Neurophysiology 94 (3), 1904 -- 1911. Lind´en, H., K. Pettersen, and G. Einevoll (2010). Intrinsic dendritic filtering gives low-pass power spectra of local field potentials. Journal of Compu- tational Neuroscience 29, 423 -- 444. Lind´en, H., T. T, T. C. Potjans, K. H. Pettersen, S. Grun, M. Diesmann, and G. T. Einevoll (2011). Modeling the spatial reach of the LFP. Neu- ron 72 (5), 859 -- 872. 29 Mainen, Z., J. Joerges, J. Huguenard, and T. Sejnowski (1995). A model of spike initiation in neocortical pyramidal neurons. Neuron 15 (6), 1427 -- 1439. Mauro, A.,F. Conti, F., F. Dodge, and R. Schor (1970). Subthreshold be- havior and phenomenological impedance of the squid giant axon. Journal of General Physiology 55 , 497 -- 532. Mazzoni, A., N. Brunel, S. Cavallari, N. K. Logothetis, and S. Panzeri (2011). Cortical dynamics during naturalistic sensory stimulations: Experiments and models. Journal of Physiology 105 (1-3), 2 -- 15. Mazzoni, A., S. Panzeri, N. K. Logothetis, and N. Brunel (2008). Encoding of naturalistic stimuli by local field potential spectra in networks of excitatory and inhibitory neurons. PLoS Computational Biology 4 (12), e1000239. Mazzoni, A., K. Whittingstall, N. Brunel, N. K. Logothetis, and S. Panz- eri (2010). Understanding the relationships between spike rate and delta/gamma frequency bands of LFPs and EEGs using a local cortical network model. NeuroImage 52 (3), 956 -- 972. Niedermeyer, E. (2005). The normal EEG of the waking adult. In E. Nieder- meyer and F. L. D. Silva (Eds.), Electroencephalography: Basic principles, clinical applications, and related fields, 5th edition, pp. 167 -- 192. Lippincott Williams & Wilkins. Nunez, P. L. and R. Srinivasan (2006). Electric Fields of the Brain: The Neurophysics of EEG (2nd ed.). New York: Oxford University Press. Omurtag, A., B. Knight, and L. Sirovich (2000). On the simulation of large populations of neurons. Journal of Computational Neuroscience 8 (1), 51 -- 63. Poulet, J.,F.A., L.M.J. Fernandez, S. Crochet, and C.H. Petersen (2012). Thalamic control of cortical states. Nature Neuroscience 15 (3), 370 -- 372. Protopapas, A., M. Vanier, and J. M. Bower (1998). Simulating large net- works of neurons. See Koch and Segev (1998), pp. 461 -- 498. Rall, W. (1977). Core conductor theory and cable properties of neurons. In E. R. Kandel (Ed.), Handbook of Physiology - The Nervous System, 30 Cellular Biology of Neurons, Volume 1, pp. 39 -- 97. American Physiological Society. Rodrigues, S., A. Chizhov, F. Marten, and J. Terry (2010). Mappings be- tween a macroscopic neural-mass model and a reduced conductance-based model. Biological Cybernetics 102 (5), 361 -- 371. Rudolph-Lilith, M., M. Dubois, and A. Destexhe (2012). Analytical integrate-and-fire neuron models with conductance-based dynamics and realistic postsynaptic potential time course for event-driven simulation strategies. Neural Computation 34, 1426 -- 1461. Sargsyan, A. R., C. Papatheodoropoulos, and G. K. Kostopoulos (2001). Modeling of evoked field potentials in hippocampal CA1 area describes their dependence on NMDA and GABA receptors. Journal of Neuroscience Methods 104, 143 -- 153. Schomer, D. L. and F. H. Lopes da Silva (Eds.) (2011). Niedermayer's Elec- troencephalography. Basic Principles, Clinical Applications, and Related Fields (6th ed.). Philadelphia: Lippincott Williams and Wilkins. Spruston, N. (2008). Pyramidal neurons: dendritic structure and synaptic integration. Nature Reviews Neuroscience 9, 206 -- 221. Wang, X.-J., J. Tegn´er, C. Constantinidis, and P. S. Goldman-Rakic (2004). Division of labor among distinct subtypes of inhibitory neurons in a cortical microcircuit of working memory. Proceedings of the National Academy of Sciences of the U.S.A. 101 (5), 1368 -- 1373. Wendling, F., J. J. Bellanger, F. Bartolomei, and P. Chauvel (2000). Rele- vance of nonlinear lumped-parameter models in the analysis of depth-EEG epileptic signals. Biological Cybernetics 83, 367 -- 378. Wilson, H. R. and J. D. Cowan (1972). Excitatory and inhibitory interactions in localized populations of model neurons. Biophysical Journal 12 (1), 1 -- 24. 31
1902.06828
2
1902
2019-07-30T08:33:59
Large-scale directed network inference with multivariate transfer entropy and hierarchical statistical testing
[ "q-bio.NC", "cs.IT", "cs.SI", "cs.IT", "physics.data-an" ]
Network inference algorithms are valuable tools for the study of large-scale neuroimaging datasets. Multivariate transfer entropy is well suited for this task, being a model-free measure that captures nonlinear and lagged dependencies between time series to infer a minimal directed network model. Greedy algorithms have been proposed to efficiently deal with high-dimensional datasets while avoiding redundant inferences and capturing synergistic effects. However, multiple statistical comparisons may inflate the false positive rate and are computationally demanding, which limited the size of previous validation studies. The algorithm we present---as implemented in the IDTxl open-source software---addresses these challenges by employing hierarchical statistical tests to control the family-wise error rate and to allow for efficient parallelisation. The method was validated on synthetic datasets involving random networks of increasing size (up to 100 nodes), for both linear and nonlinear dynamics. The performance increased with the length of the time series, reaching consistently high precision, recall, and specificity (>98% on average) for 10000 time samples. Varying the statistical significance threshold showed a more favourable precision-recall trade-off for longer time series. Both the network size and the sample size are one order of magnitude larger than previously demonstrated, showing feasibility for typical EEG and MEG experiments.
q-bio.NC
q-bio
Large-scale directed network inference with multivariate transfer entropy and hierarchical statistical testing Leonardo Novelli,1, ∗ Patricia Wollstadt,2, ∗ Pedro Mediano,3 Michael Wibral,4 and Joseph T. Lizier1 1Centre for Complex Systems, Faculty of Engineering and IT, The University of Sydney, Sydney, Australia 2Honda Research Institute Europe, Offenbach am Main, Germany 3Computational Neurodynamics Group, Department of Computing, Imperial College London, London, United Kingdom 4Campus Institute for Dynamics of Biological Networks, Georg-August University, Gottingen, Germany (Dated: July 31, 2019) Network inference algorithms are valuable tools for the study of large-scale neuroimaging datasets. Multivariate transfer entropy is well suited for this task, being a model-free measure that captures nonlinear and lagged dependencies between time series to infer a minimal directed network model. Greedy algorithms have been proposed to efficiently deal with high-dimensional datasets while avoiding redundant inferences and capturing synergistic effects. However, multiple statistical comparisons may inflate the false positive rate and are computationally demanding, which limited the size of previous validation studies. The algorithm we present -- as implemented in the IDTxl open-source software -- addresses these challenges by employing hierarchical statistical tests to control the family- wise error rate and to allow for efficient parallelisation. The method was validated on synthetic datasets involving random networks of increasing size (up to 100 nodes), for both linear and nonlinear dynamics. The performance increased with the length of the time series, reaching consistently high precision, recall, and specificity (> 98% on average) for 10 000 time samples. Varying the statistical significance threshold showed a more favourable precision-recall trade-off for longer time series. Both the network size and the sample size are one order of magnitude larger than previously demonstrated, showing feasibility for typical EEG and MEG experiments. Keywords: neuroimaging, directed connectivity, effective network, multivariate transfer entropy, information theory, nonlinear dynamics, statistical inference, nonparametric tests 9 1 0 2 l u J 0 3 ] . C N o i b - q [ 2 v 8 2 8 6 0 . 2 0 9 1 : v i X r a ∗ First authors contributed equally to this work. I. INTRODUCTION 2 The increasing availability of large-scale, fine-grained datasets provides an unprecedented opportunity for quantitative studies of complex systems. Nonetheless, a shift towards data-driven modelling of these systems requires efficient algorithms for analysing multivariate time series, which are obtained from observation of the activity of a large number of elements. In the field of neuroscience, the multivariate time series typically obtained from brain recordings serve to infer minimal (effective) network models which can explain the dynamics of the nodes in a neural system. The motivation for such models can be, for instance, to describe a causal network [1, 2] or to model the directed information flow in the system [3] in order to produce a minimal computationally equivalent network [4]. Information theory [5, 6] is well suited for the latter motivation of inferring networks that describe information flow as it provides model-free measures that can be applied at different scales and to different types of recordings. These measures, including conditional mutual information [6] and transfer entropy [7], are based purely on probability distributions and are able to identify nonlinear relationships [8]. Most importantly, information-theoretic measures allow the interpretation of the results from a distributed computation or information processing perspective, by modelling the information storage, transfer, and modification within the system [9]. Therefore, information theory simultaneously provides the tools for building the network model and the mathematical framework for its interpretation. The general approach to network model construction can be outlined as follows: for any target process (element) in the system, the inference algorithm selects the minimal set of processes that collectively contribute to the computation of the target's next state. Every process can be separately studied as a target and the results can be combined into a directed network describing the information flows in the system. This task presents several challenges: • The state space of the possible network models grows faster than exponentially with respect to the size of the network; • Information-theoretic estimators suffer from the "curse of dimensionality" for large sets of variables [10, 11]; • In a network setting, statistical significance testing requires multiple comparisons. This results in a high false positive rate (type I errors) without adequate family-wise error rate controls [12] or a high false negative rate (type II errors) with naive control procedures; • Nonparametric statistical testing based on shuffled surrogate time series is computationally demanding but currently necessary when using general information-theoretic estimators [13, 14]. Several previous studies [4, 15 -- 17] proposed greedy algorithms to tackle the first two challenges outlined above (see a summary by Bossomaier et al. [14, sec 7.2]). These algorithms mitigate the curse of dimensionality by greedily selecting the random variables that iteratively reduce the uncertainty about the present state of the target. The reduction of uncertainty is rigorously quantified by the information-theoretic measure of conditional mutual information (CMI), which can also be interpreted as a measure of conditional independence [6]. In particular, these previous studies employed multivariate forms of the transfer entropy, i.e., conditional and collective forms [18, 19]. In general, such greedy optimisation algorithms provide a locally optimal solution to the NP-hard problem of selecting the most informative set of random variables. An alternative optimisation strategy -- also based on conditional independence -- employs a preliminary step to prune the set of sources [20, 21]. Despite this progress, the computational challenges posed by the estimation of multivariate transfer entropy have severely limited the size of problems investigated in previous validation studies in the general case of nonlinear estimators, e.g., Montalto et al. [22] used 5 nodes and 512 samples; Kim et al. [23] used 6 nodes and 100 samples; Runge et al. [21] used 10 nodes and 500 samples. However, modern neural recordings often provide hundreds of nodes and tens of thousands of samples. These computational challenges, as well as the multiple testing challenges described above, are addressed here by the implementation of rigorous statistical tests, which represent the main theoretical contribution of this paper. These tests are used to control the family-wise error rate and are compatible with parallel processing, allowing the simultaneous analysis of the targets. This is a crucial feature, which enabled an improvement on the previous greedy algorithms. Exploiting the parallel computing capabilities of high-performance computing clusters and graphics processing units (GPUs) enabled the analysis of networks at a relevant scale for brain recordings -- up to 100 nodes and 10 000 samples. Our algorithm has been implemented in the recently released IDTxl Python package (the "Information Dynamics Toolkit xl" [24]).1 1 The "Information Dynamics Toolkit xl" is an open-source Python package available on GitHub (https://github.com/pwollstadt/IDTxl). In this paper, we refer to the current release at the time of writing (v1.0). We validated our method on synthetic datasets involving random structural networks of increasing size (also referred to as ground truth) and different types of dynamics (vector autoregressive processes and coupled logistic maps). In general, effective networks are able to reflect dynamic changes in the regime of the system and do not reflect an underlying structural network. Nonetheless, in the absence of hidden nodes (and other assumptions, including stationarity and the causal Markov condition), the inferred information network was proven to reflect the underlying structure for a sufficiently large sample size [17]. Experiments under these conditions provide arguably the most important validation that the algorithm performs as expected, and here we perform the first large-scale empirical validation for non-Gaussian variables. As shown in the Results, the performance of our algorithm increased with the length of the time series, reaching consistently high precision, recall, and specificity (> 98% on average) for 10 000 time samples. Varying the statistical significance threshold showed a more favourable precision-recall trade-off for longer time series. 3 II. METHODS A. Definitions and assumptions Let us consider a system of N discrete-time stochastic processes for which a finite number of samples have been recorded (over time and/or in different replications of the same experiment). In general, let us assume that the stochastic processes are stationary in each experimental time-window and Markovian with finite memory lM.2 Further assumptions will be made for the validation study. The following quantities are needed for the setup and formal treatment of the algorithm and are visualised in Figure 1 and Figure 2: target process: Y : a process of interest within the system (where Y = {Yt t ∈ N}); the choice of the target process is arbitrary and all the processes in the system can separately be studied as targets; source processes: Xi: the remaining processes within the system (where i = 1, . . . , N − 1 and Xi = {Xi,t t ∈ N}); sample number (or size): T : the number of samples recorded over time; replication number: R: the number of replications of the same experiment (e.g., trials); target present state: Yt: the random variable (RV) representing the state of the target at time t (where t ≤ T ), whose information contributors will be inferred; candidate target past: Y C i.e., Y C <t = {Yt−1, . . . , Yt−ltarget}; <t: an arbitrary finite set of RVs in the past of the target, up to a maximum lag ltarget, candidate sources past: X C <t: an arbitrary finite set of RVs in the past of the sources, up to a maximum lag lsources, i.e., X C <t = {Xi,t−1, . . . , Xi,t−lsources i = 1, . . . , N − 1}; selected target past: Y S <t: the subset of RVs within the candidate target past set Y C <t that maximally reduces the uncertainty about the present state of the target; selected sources past: X S <t that maximally further reduces the uncertainty about the present state of the target, in the context of the selected target past (explained in detail in the following section). <t: the subset of RVs within the candidate sources past set X C B. Inference algorithm For a given target process Y , the goal of the algorithm is to infer the minimal set of information contributors <t -- in the context of the relevant information contributors from the to Yt -- defined as the selected sources past X S candidate target past set, defined as the selected target past Y S <t. The algorithm operates in four steps: 2 The present state of the target does not depend on the past values of the target and the sources beyond a maximum finite lag lM. 4 FIG. 1. Example of a possible definition of the candidate sets. The bottom row represents the time series of the target process Y , with the present state Yt highlighted in green and the candidate target past set Y C <t highlighted in red (up to a lag ltarget). The remaining rows represent the time series of the source processes Xi, with the candidate sources past set X C <t highlighted in blue (up to a lag lsources). For simplicity, only a single trial of the experiment is represented. 1. Select variables in the candidate target past set Y C <t to obtain Y S <t 2. Select variables in the candidate sources past set X C <t to obtain X S <t 3. Prune the selected sources past variables 4. Test relevant variables collectively for statistical significance The operations performed in the four steps are described in detail hereafter; the result is a nonuniform embedding of the target and sources time series [15, 16, 25], as illustrated in Figure 2.3 FIG. 2. Example of a resulting nonuniform embedding of the time series relevant to Yt. The bottom row represents the time series of the target process Y , with the present state Yt highlighted in green and the selected target past set Y S <t highlighted in red (as a subset of the candidate target past set shown in light red). The remaining rows represent the time series of the source processes Xi, with the selected sources past set X S <t highlighted in blue (as a subset of the candidate sources past set shown in light blue). The embedding only specifies the relative lags between the variables. For simplicity, only a single trial of the experiment is shown. 3 The term embedding refers to the property of the selected set in capturing the underlying state of the process as it relates to the target's next value, akin to a Takens' embedding [25] yet with nonuniform delays between selected points [15, 16]. t-ltargett-1tYXN-1⋮X1t-lsourcesYtY<tCX<tCt-ltargett-1tYXN-1⋮X1t-lsourcesYtY<tSX<tS 1. Step 1: Select variables in the candidate target past set 5 The goal of the first step is to find the subset of RVs within the candidate target past set Y C <t that maximally reduces the uncertainty about the present state of the target while meeting statistical significance requirements. Let Y S <t be the selected target past set found via optimisation under these criteria. Finding the globally optimal embedding is an NP-hard problem and requires testing all the subsets of the candidate target past set. Since the number of subsets grows exponentially with the size of the candidate set, this is computationally unfeasible; therefore, a greedy approximation algorithm is employed to find a locally optimal solution in the search space of possible embeddings. This approach tackles the challenge of computational complexity by aiming at identifying a minimal conditioning set; in doing so, it also tackles the curse of dimensionality in the estimation of information-theoretic functionals. <t is initialised as an empty set and it is iteratively built up via the following algorithm: The set Y S a For each candidate variable C ∈ Y C b Find the candidate C∗ which maximises the CMI contribution (reduction of uncertainty) and perform a statistical significance test against the null hypothesis of conditional independence, i.e., that the new variable does not further reduce the uncertainty in the context of the previously included variables. If significant, add C∗ to Y S <t and remove it from Y C <t. The maximum statistic is employed to control the family-wise error rate (explained in detail in the Statistical tests section); <t, estimate the CMI contribution I(C; YtY S <t); c Repeat the previous steps until the maximum CMI contribution is not significant or Y C <t is empty. From a distributed, intrinsic computation perspective, the goal can be interpreted as finding the embedding of the target's past states that maximises the active information storage 4 [26] to ensure self-prediction optimality as suggested by Wibral et al. [27]. This approach is similar to the one proposed by Garland et al. [28] but uses nonuniform embedding and additional statistical controls. The nonuniform embedding of the time series was introduced by Vlachos and Kugiumtzis [15] and Faes et al. [16], who used an arbitrary threshold for the conditional mutual information. Lizier and Rubinov [4] introduced a statistical significance test to select the candidates, which this study builds upon in proposing the maximum statistic. In addition, they embedded the target time series before embedding the sources, i.e., the active information storage is modelled first and the information transfer is then examined in that context, thereby taking a specific modelling perspective on the information processing carried out by the system. 2. Step 2: Select variables in the candidate sources past set The goal of the second step is to find the subset of RVs within the candidate sources past set X C <t that maximally further reduces the uncertainty about the present state of the target, in the context of the selected target past found in the first step. Let X S <t be the selected sources past set found via optimisation under these criteria. As for step 1, a greedy approximation algorithm is employed and the statistical significance is tested throughout the selection process. X S a For each candidate variable C ∈ X C <t is initialised as an empty set and it is iteratively built up via the following algorithm: <t, estimate the conditional transfer entropy contribution I(C; YtY S <t) [18, 19, 29, 30]. When X S <t is empty, this is simply a pairwise or bivariate transfer entropy [7]; using the conditional form serves to prevent candidates carrying only redundant information (due to, e.g., common driver or pathway effects) from being selected, as well as to capture synergistic interactions between C and X S <t, X S <t. b Find the candidate C∗ which maximises the conditional transfer entropy contribution (reduction of uncertainty) and perform a statistical significance test against the null hypothesis of conditional independence: if significant, add C∗ to X S <t. The maximum statistic is employed to control the family-wise error rate; <t and remove it from X C c Repeat the previous steps until the maximum conditional transfer entropy contribution is not significant or X C <t is empty. From a distributed computation perspective, the goal can be interpreted as finding the nonuniform embedding of the source processes' past that maximises the collective transfer entropy to the target, defined as I(X S <t) [19]. As above, the rationale for embedding the past of the sources as a second step is to achieve optimal separation of the storage and transfer contributions [4]. <t; YtY S 4 The active information storage is defined as the mutual information between the past and the present of the target: I(Y S <t; Yt). 3. Step 3: Prune the selected sources past variables 6 The third step of the algorithm is a pruning procedure performed to ensure that the variables included in the early iterations of the second step still provide a statistically-significant information contribution in the context of the final selected sources past set X S <t. The pruning step involves the following operations: a For each variable C ∈ X S <t, X S where the the set difference operation is performed to exclude the variable C from the conditioning set; <t, estimate the conditional mutual information contribution I(C; YtY S <t \ {C}), b Find the variable C∗ which minimises the CMI contribution and perform a statistical significance test: if not <t. The minimum statistic is employed to test for significance against the null significant, remove C from X S hypothesis of conditional independence while controlling the family-wise error rate; c Repeat the previous steps until the minimum CMI contribution is not significant or X S <t is empty. The pruning step was introduced by Lizier and Rubinov [4]; remarkably, Sun et al. [17] proved that this step is essential for the theoretical convergence of the inferred network to the causal network in the Granger-Wiener framework; they also rigorously laid out the mathematical assumptions needed for such convergence (see Validation tasks section). 4. Step 4: Test relevant variables collectively for statistical significance The fourth and final step of the algorithm is the computation of the collective transfer entropy from the selected sources past set X S <t to the target and the performance of an omnibus test to ensure statistical significance against the null hypothesis of conditional independence. The resulting omnibus p-value can further be used for correction of the family-wise error rate if the inference is carried out for multiple targets. The set X S <t is only accepted as a result if all the statistical tests are passed. Importantly, the selected sources set X S <t, is the final result of the algorithm for a given target process Y . The order in which variables were inferred is not relevant. <t, inferred in the context of Y S The statistical tests play a fundamental role in the inference and provide the stopping conditions for the iterations involved in the first and second steps of the algorithm. These stopping conditions are adaptive and change according to the amount of data available (the length of the time series). Given their importance, the statistical tests are described in detail in the following section. C. Statistical tests The crucial steps in the inference algorithm rely on determining whether the CMI is positive. However, due to the finite sample size, the CMI estimators may produce non-zero estimates in the case of zero CMI and it may even return negative estimates if the estimator bias is larger than the true CMI [10, 31]. For this reason, statistical tests are required to assess the significance of the CMI estimates against the null hypothesis of no CMI (i.e., conditional independence) [3, 13, 32, 33]. For certain estimators, analytic solutions exist for the finite-sample distribution under this null hypothesis (see Lizier [34]); in the absence of an analytic solution, the null distributions are computed in a nonparametric way by using surrogate time series [35]. The surrogates are generated to satisfy the null hypothesis by destroying the temporal relationship between the source and the target while preserving the temporal dependencies within the sources. Finally, the inference algorithm is based on multiple comparisons and requires an appropriate calibration of the statistical tests to achieve the desired family-wise error rate (i.e., the probability of making one or more false discoveries, or type I errors, when performing multiple hypotheses tests). The maximum statistic and minimum statistic tests employed in this study were specifically conceived to tackle these challenges. The maximum statistic test is a step-down statistical test5 used to control the family-wise error rate when selecting the past variables for the target and sources embeddings, which involves multiple comparisons. 1. Maximum statistic test 5 A test which proceeds from the smallest to the largest p-value. When the first non-significant p-value is found, all the larger p-values are also deemed not significant. Let us first consider the first step of the main algorithm and assume that we have picked the single candidate variable C∗ (from the candidate target past set Y C <t) which maximises the CMI contribution. The maximum statistic test mirrors this selection process by picking the maximum value among the surrogates. Specifically, let I∗ := I(C∗; YtY S <t) be the maximum contribution (i.e., the maximum statistic); the following algorithm is used to test I∗ for statistical significance: 7 a For each Cj ∈ Y C j,1, . . . , C(cid:48) j,S and compute the corresponding surrogate CMI values I(cid:48) <t). More details about the surrogate generation are provided at the end of this section. The number of surrogates S must be chosen according to the desired significance level αmax, i.e., such that S > 1/αmax. <t, generate S surrogates time series C(cid:48) j,S; YtY S j,S = I(C(cid:48) j,1 = I(C(cid:48) j,1; YtY S <t), . . . , I(cid:48) b Compute the maximum CMI value over candidates I∗ Here, n denotes the number of candidates and hence the number of comparisons. The obtained values I∗ provide the (empirical) null distribution of the maximum statistic (see Table I). n,s) for each surrogate s = 1, . . . , S. 1 , . . . , I∗ S s := max(I(cid:48) 1,s, . . . , I(cid:48) c Calculate the p-value for I∗ as the fraction of surrogate maximum statistic values that are larger than I∗. d I∗ is deemed significant if the p-value is smaller than αmax (i.e., the null hypothesis of conditional independence for the candidate variable with the maximum CMI contribution is rejected at level αmax). The variables and quantities used in the above algorithm are presented in Table I. The key goal in the surrogate generation is to preserve the temporal order of samples in the target time series Yt (which is not shuffled) and preserve the distribution of the sources Cj while destroying any potential relationships between the sources and the target [3]. This can be achieved in multiple ways. If multiple replications (e.g., trials) are available, surrogate data is generated by shuffling the order of replications for the candidate Cj while keeping the order of replications for the remaining variables intact. When the number of replications is not sufficient to guarantee enough permutations, the embedded source samples within individual trials are shuffled instead (see [3, 29, 32, 33] and the summary by Lizier [34, Appendix A.5]). Note that the generation of surrogates (steps a-c) can be avoided when the null distributions can be derived analytically, e.g., with Gaussian estimators [36]. the main algorithm), with the only difference that Cj ∈ X C i.e. I(cid:48) The same test is performed during the selection of the variables in the candidate sources past set (step 2 of <t is added to the conditioning set, <t and that X S <t) for each surrogate s = 1, . . . , S. j,s; YtY S j,s = I(C(cid:48) <t, X S Variable Cj ∈ Y C CMI <t Ij = I(Cj; YtY S <t) C1 C2 ... Cn I1 I2 ... In I∗ max CMI ··· Surrogate variables Surrogate CMI 2 ··· S 1 S 2 1 1,2 ··· I(cid:48) 1,2 ··· C(cid:48) 1,1 I(cid:48) 1,S I(cid:48) C(cid:48) 1,1 C(cid:48) 2,2 ··· I(cid:48) 2,2 ··· C(cid:48) 2,1 I(cid:48) 2,S I(cid:48) 2,1 C(cid:48) C(cid:48) ... ... ... ... ... ... n,2 ··· I(cid:48) n,2 ··· C(cid:48) n,S I(cid:48) n,1 I(cid:48) C(cid:48) n,1 C(cid:48) I∗ I∗ I∗ ··· 1,S 2,S 1 2 n,S S S, obtained as I∗ TABLE I. Computing the null distribution of the maximum statistic. The null distribution is empirically described by the values I∗ 1 , . . . , I∗ n,s), for each surrogate s = 1, . . . , S. Here, n denotes the number of candidates and hence the number of comparisons. The null distribution is used to test the significance of I∗ against the null hypothesis of zero CMI. s := max(I(cid:48) 1,s, . . . , I(cid:48) 2. Family-wise error rate correction How does the maximum statistic test control the family-wise error rate? Intuitively, one or more statistics will exceed a given threshold if and only if the maximum exceeds it. This relationship can be used to obtain an adjusted threshold from the distribution of the maximum statistic under the null hypothesis, which can be used to control the family-wise error rate both in the weak and strong sense [37]. Let us quantify the false positive rate vFPR for a single variable when the maximum statistic at the significance level αmax is employed. For simplicity, the derivation is performed under the hypothesis that the information contributors to the target have been selected in the first iterations of the greedy algorithm and removed from the candidate sources past set X C <t. Under this hypothesis, the target is conditionally independent of the remaining n variables in X C <t given the selected source and target variables. Let I1, . . . , In be the corresponding CMI estimates and let Imax := max(I1, . . . , In) be the maximum statistic. As discussed above, the estimates might be positive even under the conditional independence hypothesis, due to finite-sample effects. Since the estimates are independently obtained from shuffled time series, they are treated as i.i.d. RVs. Let ithreshold be the critical threshold corresponding to the given significance value αmax, i.e., ithreshold := sup{x ∈ RP (Imax ≥ x) = αmax}. Then 8 αmax = P (Imax > ithreshold) = 1 − P (I1 ≤ ithreshold, . . . , In ≤ ithreshold) = 1 − n(cid:89) P (Ij ≤ ithreshold) Therefore, j=1 = 1 − P (I1 ≤ ithreshold)n = 1 − (1 − vFPR)n vFPR = 1 − (1 − αmax)1/n (1) (2) Interestingly, equation (2) shows that the maximum statistic correction is equivalent to the Dunn-Sid´ak correction [38]. Performing a Taylor expansion of (2) around αmax = 0 yields: − j−1(cid:81) k=0 j! ∞(cid:88) j=1 (kn − 1) (cid:16) αtarget (cid:17)j n vFPR = Truncating the Taylor series at j = 1 yields the first-order approximation vFPR ≈ αmax n , which coincides with the false positive rate resulting from the Bonferroni correction [12]. Moreover, since the summands in (3) are positive for every j, the Taylor series is lower-bounded by any truncated series. In particular, the false positive rate resulting from the Bonferroni correction is a lower bound for the vFPR (the false positive rate for a single variable resulting from the maximum statistic test), i.e., the maximum statistic correction is less stringent than the Bonferroni correction. Let us now study the effect of the maximum statistic test on the family-wise error rate tFPR for a single target while accounting for all the iterations performed during the step-down test, (i.e., tFPR is the probability that at least one of the selected sources is a false positive). We have: n(cid:88) n(cid:88) j=1 tFPR = = P ("the source selected on step j is false positive") (cid:19) (cid:18) 1 − αn max 1 − αmax αj max = αmax (3) (4) (5) (6) Therefore, j=1 tFPR ≈ αmax for the typical small values of αmax used in statistical testing (even in the limit of large n), which shows that αmax effectively controls the family-wise error rate for a single target. 3. Minimum statistic test The minimum statistic test is employed during the third main step of the algorithm (pruning step) to remove the selected variables that have become redundant in the context of the final set of selected source past variables X S <t, while controlling the family-wise error rate. This is necessary because of the multiple comparisons involved in the pruning procedure. The minimum statistic test works identically to the maximum statistic test (replacing "maximum" with "minimum" in the algorithm presented above). 4. Omnibus test 9 Let T ∗ := I(X S <t; YtY S <t) be the collective transfer entropy from all the selected sources past variables X S <t to the target Y . The value T ∗ is tested for statistical significance against the null hypothesis of zero transfer entropy (this test is referred to as the omnibus test). The null distribution is built using surrogates time series obtained via shuffling of the realisations of the selected sources (see [3, 29, 32, 33] and the summary by Lizier [34, Appendix A.5]), i.e., using a similar procedure to the one described in the Maximum statistic test section above. Testing all the selected sources collectively is in line with the perspective that the goal of the network inference is to find the set of relevant sources for each node. 5. Combining across multiple targets When the inference is performed on multiple targets, the omnibus p-values can be employed in further statistical tests to control the family-wise error rate for the overall network (e.g., via FDR-correction [12, 39], which is implemented in the IDTxl toolbox). It is important to fully understand the statistical questions and validation procedure implied by this approach. Combining the results across multiple targets by reusing the omnibus test p-values for the FDR-correction yields a hierarchical test. The test answers two nested questions: (1) 'which nodes receive any significant overall information transfer?' and, if any, (2)'what is the structure of the incoming information transfer to each node?'. However, the answers are computed in the reverse order, for the following reason: it would be computationally unfeasible to directly compute the collective transfer entropy from all candidate sources to the target right at the beginning of the network inference process. At this point, the candidate source set usually contains a large number of variables so that estimation will likely fall prey to the curse of dimensionality. Instead, a conservative approximation of the collective information transfer is obtained by considering only a subset of the potential sources, i.e., those deemed significant by the maximum and minimum statistic tests described in the previous sections. Only if this approximation of the total information transfer is also deemed significant by the omnibus test (as well as by the FDR test at the network level), then the subset of significant sources for that target is interpreted post-hoc as the local structure of the incoming information transfer. This way, the testing procedure exhibits a hierarchical structure: the omnibus test operates at the higher (global) level concerned with the collective information transfer, whereas the minimum and maximum tests operate at the lower (local) level of individual source-target connections. Compared to a non-hierarchical analysis with a correction for multiple comparisons across all links (e.g., by network- wide Bonferroni correction or the use of the maximum statistic across all potential links), the above strategy buys both statistical sensitivity ('recall') and the possibility to trivially parallelise computations across targets. The price to be paid is that a link with a relatively strong information transfer into a node with non-significant overall incoming information transfer may get pruned, while a link with relatively weaker information transfer into a node with significant overall incoming information transfer will prevail. This behaviour clearly differs from a correction for multiple comparisons across all links. Arguably, this difference is irrelevant in many practical cases, although it could become noticeable for networks with high average in-degree and relatively uniform information transfer across the links. The difference can be reduced by setting a conservative critical threshold for the lower-level greedy analysis. D. Validation tasks For the purpose of the validation study, the additional assumptions of causal sufficiency 6 and the causal Markov condition 7 were made, such that the inferred network was expected to closely reflect the structural network for a sufficiently large sample size [17]. Although this is not always the case, experiments under these conditions allow the evaluation of the performance of the algorithm with respect to an expected ground truth. An intuitive definition of these conditions is provided here, while the technical details are discussed at length in [40]. Moreover, the intrinsic stochastic nature of the processes makes purely synergistic and purely redundant interactions unlikely (and indeed vanishing for large sample size), thus satisfying the faithfulness condition [40]. The complete network inference algorithm implemented in the IDTxl toolkit (release v1.0) was validated on multiple synthetic datasets, where both the structural connectivity and the dynamics were known. Given the general scope of the 6 Causal sufficiency: The set of observed variables includes all their common causes (or the unobserved common causes have constant values). 7 Causal Markov condition: A variable X is independent of every other past variable conditional on all of its direct causes. 10 toolkit, two dynamical models of broad applicability were chosen: a vector autoregressive process (VAR) and a coupled logistic maps process (CLM); both models are widely used in computational neuroscience [41 -- 43], macroeconomics [44, 45], and chaos theory [46]. The primary goal was to quantify the scaling of the performance with respect to the size of the network and the length of the time series. Sparse directed random Erdos-R´enyi networks [47] of increasing size (N = 10 to 100 nodes) were generated with a link probability p = 3/N to obtain an expected in-degree of 3 links. Both the VAR and the CLM stochastic processes were repeatedly simulated on each causal network with increasingly longer time series (T = 100 to 10 000 samples), a single replication (or trial, i.e., R = 1), and with 10 random initial conditions. The performance was evaluated in terms of precision, recall, and specificity in the classification of the links. Further simulations were carried out to investigate the influence of the critical alpha level for statistical significance and the performance of different estimators of conditional mutual information. 1. Vector autoregressive process The specific VAR process used in this study is described by the following discrete-time recurrence relation: Yt = βYt−1 + αX Xt−lX + ηt (7) (cid:80) where XY denotes the set of causal sources of the target process Y and a single random lag lX ∈ {1, 2, 3, 4, 5} was used for each source X ∈ XY . A Gaussian noise term ηt with mean µ = 0 and standard deviation θ = 0.1 was added at each time step t; the noise terms added to different variables were uncorrelated. The self-coupling coefficient was set to β = 0.5 and the cross-coupling coefficients αX were uniform and normalised for each target such that αX = 0.4. This choice of parameters guaranteed that the VAR processes were stable (the resulting spectral radii were between 0.9 and 0.95) and had stationary multivariate Gaussian distributions [48]. As such, the Gaussian estimator implemented in IDTxl was employed for transfer entropy measurements in VAR processes. Note that transfer entropy and Granger causality [49] are equivalent for Gaussian variables [50]; therefore, using the Gaussian estimator with our algorithm can be viewed as extending Granger causality in the same multivariate/greedy fashion. X∈XY 2. Coupled logistic maps process The coupled logistic maps process used in this study is described by the following discrete-time recurrence relations: (cid:88) X∈XY (cid:88) at = βYt−1 + αX Xt−lX Yt = (4at(1 − at) + ηt) mod 1 X∈XY (8) At each time step t, each node Y computes the weighted input at as a linear combination of its past value and the past of its sources, with the same conditions used for the VAR process on the choice of the random lags lX and coupling coefficients β and αX . The value Yt is then computed by applying the logistic map activation function f (x) = 4x(1− x) to the weighted input at and adding the Gaussian noise ηt with the same properties used for the VAR process. Notice that the coefficient (r = 4) used in the logistic map function corresponds to the fully-developed chaotic regime. The modulo-1 operation ensures that Yt ∈ [0, 1] after the addition of noise. The nearest-neighbour estimators were employed for transfer entropy measurements in the analysis of the CLM processes (in particular, Kraskov's estimator I (1) with k = 4 nearest neighbours [31] and its extension to CMI [51 -- 53]). Nearest-neighbour estimators are model-free and are able to detect nonlinear dependencies in stochastic processes with non-Gaussian stationary distributions; fast CPU and GPU implementations are provided by the IDTxl package. III. RESULTS A. Influence of network size and length of the time series The aim of the first analysis was to quantify the scaling of the performance with respect to the size of the network and the length of the time series. 11 The inferred network was built by adding a directed link from a source node X to a target node Y whenever a significant transfer entropy from X to Y was measured while building the selected sources past set X S <t (i.e., whenever X ∩ X S <t (cid:54)= ∅). The critical alpha level for statistical significance was set to αmax = 0.001 and S = 1000 surrogates were used for all experiments unless otherwise stated. The candidate sets for the target as well as the sources were initialised with a maximum lag of five (i.e., ltarget = lsources = 5, corresponding to the largest lag values used in the definition of the VAR and CLM processes). The network inference performance was evaluated in comparison to the known underlying structural network as a binary classification task, using standard statistics based on the number of true positives (TP, i.e., correctly classified existing links), false positives (FP, i.e., absent links falsely classified as existing), true negatives (TN, i.e., correctly classified absent links), and false negatives (FN, i.e., existing links falsely classified as absent). The following standard statistics were employed in the evaluation: precision: = T P/(T P + F P ) recall: = T P/(T P + F N ) specificity: = T N/(T N + F P ) The plots in Figure 3 summarise the results in terms of precision and recall, while the specificity is additionally plotted in the Supplementary materials. For both types of dynamics, the performance increased with the number of samples and decreased with the size of the network. For shorter time series (T = 100 and T = 1000), the recall was the most affected performance measure as a function of N and T , while the precision and the specificity were always close to optimal (> 98% on average). (Note that, while S = 1000 is minimal for αmax = 0.001, recall was unchanged using S = 10 000 for T = 100). For longer time series (T = 10 000), high performance according to all measures was achieved for both the VAR and CLM processes, regardless of the size of the network. The high precision and specificity are due to the effective control of the false positives, in accordance with the strict statistical significance level αmax = 0.001 (the influence of αmax is further discussed in the following sections). The inference algorithm was therefore conservative in the classification of the links. VAR CLM FIG. 3. Precision (top) and recall (bottom) for different network sizes, sample sizes, and dynamics. Left: Vector autoregressive process; Right: Coupled logistic maps. Each subplot shows five curves, corresponding to different time series lengths (T = 100, 300, 1000, 3000, 10 000). The results for 10 simulations from different initial conditions are shown (low-opacity markers) in addition to the mean values (solid markers). All the random networks have an average in-degree N p = 3. 20406080100Networksize0.00.20.40.60.81.0PrecisionT=100T=300T=1000T=3000T=100001040701000.900.951.0020406080100Networksize0.00.20.40.60.81.0PrecisionT=100T=300T=1000T=3000T=100001040701000.900.951.0020406080100Networksize0.00.20.40.60.81.0RecallT=100T=300T=1000T=3000T=1000020406080100Networksize0.00.20.40.60.81.0RecallT=100T=300T=1000T=3000T=10000 B. Validation of false positive rate 12 The critical alpha level for statistical significance αmax is a parameter of the algorithm which is designed to control the number o false positives in the network inference. As discussed in the Statistical tests section in the Methods, αmax controls the probability that a target is a false positive, i.e., that at least one of its sources is a false positive. This approach is in line with the perspective that the goal of the network inference is to find the set of relevant sources for each node. A validation study was carried out to verify that the final number of false positives is consistent with the desired level αmax after multiple statistical tests are performed. The false positive rate (i.e. F P/(F P + T N )) was computed after performing the inference on empty networks, where every inferred link is a false positive by definition (i.e., under the complete null hypothesis). The rate was in good accordance with the critical alpha threshold αmax for all network sizes, as shown in Figure 4. The false positive rate validation was replicated in a scenario where the null hypothesis held for real fMRI data from the Human Connectome Project resting-state dataset (see Supporting Information). The findings are presented in the Supplementary Material, together with a note on autocorrelation. Notably, the results on fMRI data are in agreement with the results on synthetic data shown in Figure 4. FIG. 4. Validation of false positive rate for a single target (tFPR) on empty networks. The points indicate the average false positive rate over 50 simulations of a vector autoregressive process (T = 10 000). The horizontal marks indicate the corresponding 5th and 95th percentiles of the expected range. These were computed empirically from the distribution of the random variable (cid:104)Xj/N(cid:105), where Xj ∼ Binomial(N, αmax) are i.i.d. random variables, and the angular brackets indicate the finite average over 50 repetitions. The 5th percentile for N = 10 and N = 40 and αmax = 10−3 are equal to zero and therefore omitted from the log-log plot. The identity function is plotted as a reference (dashed line). C. Influence of critical level for statistical significance Given the conservative results obtained for both the VAR and CLM processes (Figure 3), a natural question is to what extent the recall could be improved by increasing the critical alpha level αmax and to what extent the precision would be negatively affected as a side effect. In order to elucidate this trade-off, the analysis described above (Figure 3) was repeated for increasing values of αmax, with results shown in Figure 5. For the shortest time series (T = 100), increasing αmax resulted in a higher recall and a lower precision, as expected; on the other hand, for the longest time series (T = 10 000), the performance measures were not significantly affected. Interestingly, for the intermediate case (T = 1000), increasing αmax resulted in higher recall without negatively affecting the precision. 103102max103102FalsepositiverateN=10N=40N=70N=100Identity 13 VAR CLM FIG. 5. Influence of statistical significance threshold on network inference performance. Precision vs. recall for different statistical significance levels (αmax = 0.05, 0.01, 0.001), corresponding to different colours. The plots summarise the results for different dynamics (Top: Vector autoregressive process; Bottom: Coupled logistic maps), different time series lengths (T = 100, 1000, 10 000), and different network sizes (N = 10, 40, 70, 100, not distinguished). The arrows join the mean population values for the lowest and highest significance levels, illustrating the average trade-off between precision loss and recall gain. D. Inference of coupling lags So far, the performance evaluation focused on the identification the correct set of sources for each target node, regardless of the coupling lags. However, since the identification of the correct coupling lags is particularly relevant in neuroscience (see Wibral et al. [27] and references therein), the performance of the algorithm in identifying the correct coupling lags was additionally investigated. By construction, a single coupling lag was imposed between each pair of processes (chosen at random between 0.00.20.40.60.81.0Recall0.60.81.0PrecisionT=100T=1000T=10000max=0.001max=0.01max=0.050.00.20.40.60.81.0Recall0.50.60.70.80.91.0PrecisionT=100T=1000T=10000max=0.001max=0.01max=0.05 one and five discrete time steps, as described in the Methods). The average absolute error between the real and the inferred coupling lags was computed on the correctly recalled sources and divided by the value expected at random (which is the average absolute difference between two i.i.d. random integers in the [1, 5] interval). In line with the previous results on precision, the absolute error on coupling lag is consistently much smaller than that expected at random, even for the shortest time series (Figure 6). Furthermore, 1000 samples were sufficient to achieve nearly optimal performance for both the VAR and the CLM processes, regardless of the size of the network. Note that as T increases and the recall increases, the lag error can increase (c.f. T = 100 to 300 for the CLM process). This is perhaps because while the larger T permits more weakly contributing sources to be identified, it is not large enough to reduce the estimation error to make lag identification on these sources precise. VAR CLM 14 FIG. 6. Average absolute error between the real and the inferred coupling lags, relative to the value expected at random. Results for different dynamics (Left: Vector autoregressive process; Right: Coupled logistic maps), different time series lengths (T = 100, 300, 1000, 3000, 10 000), and different network sizes (N = 10, 40, 70, 100). The error bars indicate the standard deviation over 10 simulations from different initial conditions. E. Estimators Given its speed, the Gaussian estimator is often used for large datasets or as a first exploratory step, even when the stationary distribution cannot be assumed to be Gaussian. The availability of the ground truth allowed to compare the performance of the Gaussian estimator and the nearest-neighbour estimator on the nonlinear CLM process, which does not satisfy the Gaussian assumption. As expected, the performance of the Gaussian estimator was lower than the performance of the nearest-neighbour estimator for all network sizes (Figure 7). FIG. 7. Gaussian vs. nearest-neighbour estimator on the coupled logistic maps process. The precision (left) and recall (right) are plotted against the network size and a fixed time series length (T = 10 000 samples). The results for 10 simulations from different initial conditions are shown (low-opacity markers) in addition to the mean values (solid markers). The statistical significance level αmax = 0.05 was employed; an even larger gap between the recall of the estimators is obtained with αmax = 0.001. The hierarchical tests introduced in the Methods section allow running the network inference algorithm in parallel on a high-performance computing cluster. Such parallelisation is especially needed when employing the nearest-neighbour estimator. In particular, each target node can be analysed in parallel on a CPU (employing one or more cores) or a GPU, which is made possible by the CPU and GPU estimators provided by the IDTxl package (custom OpenCL 102103104T0.00.10.20.3LagerrorN=10N=40N=70N=100102103104T0.00.10.20.3LagerrorN=10N=40N=70N=10020406080100Networksize0.00.51.0PrecisionNearest-neighbour estimatorGaussian estimator1040701000.940.971.0020406080100Networksize0.00.51.0RecallNearest-neighbour estimatorGaussian estimator kernels were written for the GPU implementation). A summary of the CPU and GPU run-times is provided in the Supplementary materials. 15 IV. DISCUSSION The algorithm presented in this paper provides robust statistical tests for network inference to control the false positive rate. These tests are compatible with parallel computation on high-performance computing clusters, which enabled the validation study on synthetic sparse networks of increasing size (10 to 100 nodes), using different dynamics (linear autoregressive processes and nonlinear coupled logistic maps) and increasingly longer time series (100 to 10 000 samples). Both the network size and the sample size are one order of magnitude larger than previously demonstrated, showing feasibility for typical EEG and MEG experiments. The results demonstrate that the statistical tests achieve the desired false positive rate and successfully address the multiple-comparison problems inherent in network inference tasks (Figure 4). The ability to control the false positives while building connectomes is a crucial prerequisite for the application of complex network measures, to the extent that [54] concluded that "specificity is at least twice as important as sensitivity [i.e., recall] when estimating key properties of brain networks, including topological measures of network clustering, network efficiency and network modularity". The reason is that false positives occur more prevalently between network modules than within them and the spurious inter-modular connections have a dramatic impact on network topology [54]. The trade-off between precision and recall when relaxing the statistical significance threshold was further investigated (Figure 5). When only 100 samples were used, the average recall gain was more than five times smaller than the average precision loss. In our opinion, this result is possibly due to the sparsity of the networks used in this study and suggests a conservative choice of the threshold for sparse networks and short time series. The trade-off was reversed for longer time series: when 1000 samples were used, the average recall gain was more than five times larger than the average precision loss. Finally, for 10 000 samples, high precision and recall were achieved (> 98% on average) for both the vector autoregressive and the coupled logistic maps processes, regardless of the statistical significance threshold. For both types of dynamics, the network inference performance increased with the length of the time series and decreased with the size of the network (Figure 3). This is to be expected since larger systems require more statistical tests and hence stricter conditions to control the family-wise error rate (false positives). Specifically, larger networks result in wider null distributions of the maximum statistic (i.e., larger variance), whereas longer time series have the opposite effect. Therefore, for large networks and short time series, controlling the false positives can have a negative impact on the ability to identify the true positives, particularly when the effect size (i.e., the transfer entropy value) is small. In addition, the superior ability of the nearest-neighbour estimator over the Gaussian estimator in detecting nonlinear dependencies was quantified. There is a critical motivation for this comparison: the general applicability of the nearest-neighbour estimators comes at the price of higher computational complexity and a significantly longer run-time, so that the Gaussian estimator is often used for large datasets (or at least as a first exploratory step), even when the Gaussian hypothesis is not justified. To investigate such scenario, the Gaussian estimator was tested on the nonlinear logistic map processes: while the resulting recall was significantly lower than the nearest-neighbour estimator for all network sizes, it was nonetheless able to identify over half of the links for a sufficiently large number (10 000) of time samples (Figure 7). The stationarity assumption about the time series corresponds to assuming a single regime of neuronal activity in real brain recordings. If multiple regimes are recorded, which is typical in experimental settings (e.g., sequences of tasks or repeated presentation of stimuli interleaved with resting time windows), different stationary regimes can be studied by performing the analysis within each time window. The networks obtained in different time windows can either be studied separately and compared against each other or collectively interpreted as a single evolving temporal network. To obtain a sufficient amount of observations per window, multiple replications of the experiment under the same conditions are typically carried out. Replications can be assumed to be cyclo-stationary and estimation techniques exploiting this property have been proposed [53, 55]; these estimators are also available in the IDTxl Python package. The convergence to the (unknown) causal network was only proven under the hypotheses of stationarity, causal sufficiency, and the causal Markov condition [17]. However, conditional independence holds under milder assumptions [56] and the absence of links is valid under general conditions. The conditional independence relationships can, therefore, be used to exclude variables in following intervention-based causal experiments, making network inference methods valuable for exploratory studies. In fact, the directed network is only one part of the model and provides the scaffold over which the information- theoretic measures are computed. Therefore, even if the structure of a system is known and there is no need for network inference, information theory can still provide nontrivial insights on the distributed computation by modelling 16 the information storage, transfer, and modification within the system [9]. This decomposition of the predictive information into the active information storage and transfer entropy components is one out of many alternatives within the framework proposed by Chicharro and Ledberg [57]. Arguably, the storage-transfer decomposition reflects the segregation-integration dichotomy that has long characterised the interpretation of brain function [58, 59]. Information theory has the potential to provide a quantitative definition of these fundamental but still unsettled concepts [60]. In addition, information theory provides a new way of testing fundamental computational theories in neuroscience, e.g., predictive coding [61]. As such, information-theoretic methods should not be seen as opposed to model-based approaches, but complementary to them [62]. If certain physically motivated parametric models are assumed, the two approaches are equivalent for network inference: maximising the log-likelihood is asymptotically equivalent to maximising the transfer entropy [36, 63]. Moreover, different approaches can be combined, e.g., the recent large-scale application of spectral DCM was made possible by using functional connectivity models to place prior constraints on the parameter space [64]. Networks inferred using bivariate transfer entropy have also been employed to reduce the model space prior to DCM analysis [65]. In conclusion, the continuous evolution and combination of methods show that network inference from time series is an active field of research and there is a current trend of larger validation studies, statistical significance improvements, and reduction of computational complexity. Information-theoretic approaches require efficient tools to employ nearest- neighbour estimators on large datasets of continuous-valued time series, which are ubiquitous in large-scale brain recordings (calcium imaging, EEG, MEG, fMRI). The algorithm presented in this paper is compatible with parallel computation on high-performance computing clusters, which enabled the study of synthetic nonlinear systems of 100 nodes and 10 000 samples. Both the network size and the sample size are one order of magnitude larger than previously demonstrated, bringing typical EEG and MEG experiments into scope for future information-theoretic network inference studies. Furthermore, the statistical tests presented in the Methods are generic and compatible with any underlying conditional mutual information or transfer entropy estimators, meaning that estimators applicable to spike trains [66] can be used with this algorithm in future studies. V. SUPPORTING INFORMATION The network inference algorithm described in this paper is implemented in the open-source Python software package IDTxl [24], which is freely available on GitHub (https://github.com/pwollstadt/IDTxl). In this paper, we refer to the current release (v1.0) at the time of writing (doi:10.5281/zenodo.2554339). The raw data used for the experiment presented in the Supplementary Material is openly available on the MGH-USC Human Connectome Project database (https://ida.loni.usc.edu/login.jsp). VI. ACKNOWLEDGEMENTS This research was supported by: Universities Australia/German Academic Exchange Service (DAAD) Australia- Germany Joint Research Cooperation Scheme grant: "Measuring neural information synthesis and its impairment", Grant/Award Number: PPP Australia Projekt-ID 57216857; Deutsche Forschungsgemeinschaft (DFG) Grant CRC 1193 C04; and Australian Research Council DECRA grant DE160100630 and Discovery grant DP160102742. Data collection and sharing for the experiment presented in the Supplementary material was provided by the MGH-USC Human Connectome Project (HCP; Principal Investigators: Bruce Rosen, M.D., Ph.D., Arthur W. Toga, Ph.D., Van J. Weeden, MD). HCP funding was provided by the National Institute of Dental and Craniofacial Research (NIDCR), the National Institute of Mental Health (NIMH), and the National Institute of Neurological Disorders and Stroke (NINDS). HCP data are disseminated by the Laboratory of Neuro Imaging at the University of Southern California. The authors acknowledge the Sydney Informatics Hub and the University of Sydney's high-performance computing cluster Artemis for providing the high-performance computing resources that have contributed to the research results reported within this paper. Furthermore, the authors thank Aaron J. Gutknecht for commenting on a draft of this paper, and Oliver Cliff for useful discussions and comments. VII. AUTHOR CONTRIBUTIONS Leonardo Novelli: Conceptualization; Data curation; Formal analysis; Investigation; Software; Validation; Visu- alization; Writing -- original draft. Patricia Wollstadt: Conceptualization; Software; Writing -- review & editing. Pedro Mediano: Software; Writing -- review & editing. Michael Wibral: Conceptualization; Funding acquisition; Methodology; Software; Supervision; Writing -- review & editing. Joseph T. Lizier: Conceptualization; Funding acquisition; Methodology; Software; Supervision; Writing -- review & editing. 17 [1] K. J. Friston, Human Brain Mapping 2, 56 (1994). [2] N. Ay and D. Polani, Advances in Complex Systems 11, 17 (2008). [3] R. Vicente, M. Wibral, M. Lindner, and G. Pipa, Journal of Computational Neuroscience 30, 45 (2011). [4] J. T. Lizier and M. Rubinov, Multivariate construction of effective computational networks from observational data, Tech. Rep. Preprint 25/2012 (Max Planck Institute for Mathematics in the Sciences, 2012). [5] C. E. Shannon, Bell System Technical Journal 27, 379 (1948). [6] T. M. Cover and J. A. Thomas, Elements of Information Theory, 2nd ed. (John Wiley & Sons, Inc., Hoboken, NJ, USA, 2005) p. 748. [7] T. Schreiber, Physical Review Letters 85, 461 (2000). [8] M. Palus, V. Albrecht, and I. Dvo´ak, Physics Letters A 175, 203 (1993). [9] J. T. Lizier, The Local Information Dynamics of Distributed Computation in Complex Systems, Springer Theses (Springer Berlin Heidelberg, Berlin, Heidelberg, 2013) p. 311. [10] M. S. Roulston, Physica D: Nonlinear Phenomena 125, 285 (1999). [11] L. Paninski, Neural Computation 15, 1191 (2003). [12] T. Dickhaus, Simultaneous Statistical Inference (Springer Berlin Heidelberg, Berlin, Heidelberg, 2014) p. 180. [13] M. Lindner, R. Vicente, V. Priesemann, and M. Wibral, BMC Neuroscience 12, 119 (2011). [14] T. Bossomaier, L. Barnett, M. Harr´e, and J. T. Lizier, An Introduction to Transfer Entropy (Springer International Publishing, Cham, 2016) p. 190. [15] I. Vlachos and D. Kugiumtzis, Physical Review E 82, 016207 (2010). [16] L. Faes, G. Nollo, and A. Porta, Physical Review E 83, 051112 (2011). [17] J. Sun, D. Taylor, and E. M. Bollt, SIAM Journal on Applied Dynamical Systems 14, 73 (2015). [18] J. T. Lizier, M. Prokopenko, and A. Y. Zomaya, Physical Review E 77, 026110 (2008). [19] J. T. Lizier, M. Prokopenko, and A. Y. Zomaya, Chaos 20, 037109 (2010). [20] J. Runge, J. Heitzig, V. Petoukhov, and J. Kurths, Physical Review Letters 108, 258701 (2012). [21] J. Runge, P. Nowack, M. Kretschmer, S. Flaxman, and D. Sejdinovic, arXiv Preprint (2018), arXiv: 1702.07007. [22] A. Montalto, L. Faes, and D. Marinazzo, PLoS ONE 9, e109462 (2014). [23] P. Kim, J. Rogers, J. Sun, and E. M. Bollt, Journal of Computational and Nonlinear Dynamics 12, 011008 (2016). [24] P. Wollstadt, J. T. Lizier, R. Vicente, C. Finn, M. Mart´ınez-Zarzuela, P. Mediano, L. Novelli, and M. Wibral, Journal of Open Source Software 4, 1081 (2019). [25] F. Takens, in Dynamical Systems and Turbulence, edited by D. Rand and L. Young (Springer Berlin Heidelberg, 1981) pp. 366 -- 381. [26] J. T. Lizier, M. Prokopenko, and A. Y. Zomaya, Information Sciences 208, 39 (2012). [27] M. Wibral, N. Pampu, V. Priesemann, F. Siebenhuhner, H. Seiwert, M. Lindner, J. T. Lizier, and R. Vicente, PLoS ONE 8, e55809 (2013). [28] J. Garland, R. G. James, and E. Bradley, Physical Review E 93, 022221 (2016). [29] P. F. Verdes, Physical Review E 72, 026222 (2005). [30] V. A. Vakorin, O. A. Krakovska, and A. R. McIntosh, Journal of Neuroscience Methods 184, 152 (2009). [31] A. Kraskov, H. Stogbauer, and P. Grassberger, Physical Review E 69, 066138 (2004). [32] M. Ch´avez, J. Martinerie, and M. Le Van Quyen, Journal of Neuroscience Methods 124, 113 (2003). [33] J. T. Lizier, J. Heinzle, A. Horstmann, J.-D. Haynes, and M. Prokopenko, Journal of Computational Neuroscience 30, 85 (2011). [34] J. T. Lizier, Frontiers in Robotics and AI 1, 11 (2014). [35] T. Schreiber and A. Schmitz, Physica D: Nonlinear Phenomena 142, 346 (2000). [36] L. Barnett and T. Bossomaier, Physical Review Letters 109, 138105 (2012). [37] T. Nichols and S. Hayasaka, Statistical Methods in Medical Research 12, 419 (2003). [38] Z. Sid´ak, Journal of the American Statistical Association 62, 626 (1967). [39] Y. Benjamini and Y. Hochberg, Journal of the Royal Statistical Society. Series B (Methodological) 57, 289 (1995). [40] P. Spirtes, C. Glymour, and R. Scheines, Causation, Prediction, and Search, Lecture Notes in Statistics, Vol. 81 (Springer New York, 1993). [41] A. Zalesky, A. Fornito, L. Cocchi, L. L. Gollo, and M. Breakspear, Proceedings of the National Academy of Sciences 111, 10341 (2014). [42] M. Rubinov, O. Sporns, C. van Leeuwen, and M. Breakspear, BMC Neuroscience 10, 55 (2009). [43] P. A. Valdes-Sosa, A. Roebroeck, J. Daunizeau, and K. J. Friston, NeuroImage 58, 339 (2011). [44] C. A. Sims, Econometrica 48, 1 (1980). [45] H.-W. Lorenz, in Nonlinear Dynamical Economics and Chaotic Motion (Springer Berlin Heidelberg, 1993) pp. 119 -- 166. [46] S. H. Strogatz, Nonlinear Dynamics and Chaos (CRC Press, Boca Raton, 2015) p. 531. 18 [47] P. Erdos and A. R´enyi, Publicationes Mathematicae Debrecen 6, 290 (1959). [48] F. M. Atay and O. Karabacak, SIAM Journal on Applied Dynamical Systems 5, 508 (2006). [49] C. W. J. Granger, Econometrica 37, 424 (1969). [50] L. Barnett, A. B. Barrett, and A. K. Seth, Physical Review Letters 103, 238701 (2009). [51] S. Frenzel and B. Pompe, Physical Review Letters 99, 204101 (2007). [52] M. Vejmelka and M. Palus, Physical Review E 77, 026214 (2008). [53] G. G´omez-Herrero, W. Wu, K. Rutanen, M. Soriano, G. Pipa, and R. Vicente, Entropy 17, 1958 (2015). [54] A. Zalesky, A. Fornito, L. Cocchi, L. L. Gollo, M. P. van den Heuvel, and M. Breakspear, NeuroImage 142, 407 (2016). [55] P. Wollstadt, M. Mart´ınez-Zarzuela, R. Vicente, F. J. D´ıaz-Pernas, and M. Wibral, PLoS ONE 9, e102833 (2014). [56] J. Runge, Chaos 28, 075310 (2018). [57] D. Chicharro and A. Ledberg, Physical Review E 86, 041901 (2012). [58] S. Zeki and S. Shipp, Nature 335, 311 (1988). [59] O. Sporns, Networks of the Brain (MIT Press, 2010) p. 424. [60] M. Li, Y. Han, M. J. Aburn, M. Breakspear, R. A. Poldrack, J. M. Shine, and J. T. Lizier, bioRxiv Preprint , 581538 (2019). [61] A. Brodski-Guerniero, G.-F. Paasch, P. Wollstadt, I. Ozdemir, J. T. Lizier, and M. Wibral, The Journal of Neuroscience 37, 8273 (2017). [62] K. J. Friston, R. Moran, and A. K. Seth, Current Opinion in Neurobiology 23, 172 (2013). [63] O. Cliff, M. Prokopenko, and R. Fitch, Entropy 20, 51 (2018). [64] A. Razi, M. L. Seghier, Y. Zhou, P. McColgan, P. Zeidman, H.-J. Park, O. Sporns, G. Rees, and K. J. Friston, Network Neuroscience 1, 222 (2017). [65] J. S. Chan, M. Wibral, P. Wollstadt, C. Stawowsky, M. Brandl, S. Helbling, M. Naumer, and J. Kaiser, bioRxiv Preprint , 178095 (2017). [66] R. E. Spinney, M. Prokopenko, and J. T. Lizier, Physical Review E 95, 032319 (2017). [67] D. Van Essen, K. Ugurbil, E. Auerbach, D. Barch, T. Behrens, R. Bucholz, A. Chang, L. Chen, M. Corbetta, S. Curtiss, S. Della Penna, D. Feinberg, M. Glasser, N. Harel, A. Heath, L. Larson-Prior, D. Marcus, G. Michalareas, S. Moeller, R. Oostenveld, S. Petersen, F. Prior, B. Schlaggar, S. Smith, A. Snyder, J. Xu, and E. Yacoub, NeuroImage 62, 2222 (2012). [68] L. Barnett and A. K. Seth, Journal of Neuroscience Methods 201, 404 (2011). [69] J. Theiler, Physical Review A 34, 2427 (1986). [70] H. Kantz and T. Schreiber, Nonlinear Time Series Analysis, 2nd ed. (Cambridge University Press, 2003). VIII. SUPPORTING INFORMATION A. Run-time on CPU and GPU 19 The number of transfer entropy calculations scales as O(N 2dlmaxS), where N is the number of processes, d is the average inferred in-degree, lmax is the maximum temporal search depth per process (i.e., lmax = max{ltarget, lsources}), and S is the number of surrogates. This assumes that d is independent of the network size N ; however, note that d = N in the worst case of a fully connected network, leading to cubic run-times. The network inference algorithm was designed for parallelisation: • When using the CPU estimators, it is possible to parallelise over targets, resulting in O(N dlmaxS) transfer entropy calculations per target for the nearest-neighbour estimator and O(N dlmax) transfer entropy calculations for the Gaussian estimator (if using analytic null distributions instead of surrogates). The complexity of each calculation is O(kT logT ) for the nearest-neighbour estimator and O(T ) for the Gaussian estimator (where T is number of time series samples and k is the number of nearest-neighbours). • When using the GPU estimators, it is possible to parallelise both over targets and surrogates. Each target requires O(N dlmax) transfer entropy calculations including surrogates, assuming that all surrogates for a source fit into the GPU's main memory and can be processed in parallel. The complexity of each calculation is O(T 2). If enough memory is available, it is further possible to parallelise over time samples T , resulting in faster run-times in practice. The practical run-time for a full network analysis that considers each process as a target depends on the number of available computing nodes. In the worst case, where only a single computing node is available, the full run-time is equal to the single-target run-time multiplied by N , since the target are analysed in series. In the best case, if N computing nodes are available, the full run-time is equal to the single-target run-time, since all targets can be analysed in parallel. Notice that there is a trade-off between run-time and memory requirements: if all the targets are analysed in parallel, the full required memory in N times larger than the memory required in the single-node case; conversely, if the targets are analysed in series, the full required memory is equal to the memory required in the single-node case. In the experiments presented in this article, the algorithm was either run using a single core per target (on different Intel Xeon CPUs with similar characteristics: 2.1-2.6 GHz), or using a whole dedicated GPU per target (NVIDIA V100 SXM2, 16 GB RAM). These computations were performed on the Artemis computing cluster made available by the Sydney Informatics Hub at The University of Sydney. The maximum CPU and GPU run-times for a single target are shown in Table II, which summarises the results for different time series lengths and different network sizes. Notice that the CPU run-time per target can be reduced if multiple cores per target are available. Sample size Network size max GPU time max CPU time max CPU memory per target per target per target T 100 1000 10 000 N 10 40 70 100 10 40 70 100 10 40 70 100 (h) 0.005 0.01 0.02 0.03 0.05 0.20 0.25 0.45 0.85 3.30 5.00 6.70 (h) 0.01 0.05 0.10 0.15 0.63 2.30 4.30 5.50 13.00 120.00 250.00 335.00 (GB) <0.50 <0.50 <0.50 0.63 0.65 0.70 0.70 0.70 5.50 6.60 7.40 8.00 TABLE II. Maximum CPU and GPU run-time for a single target using the nearest-neighbour estimator and 200 surrogates. Summary of the results for different time series lengths (T = 100, 1000, 10 000) and different network sizes (N = 10, 40, 70, 100). 20 B. Validation of false positive rate on real fMRI data The false positive rate validation (presented in Figure 4 for synthetic VAR data) was replicated in a scenario where the null hypothesis held for real data. Once again, the aim was to verify that the false positive rate was consistent with the desired level αmax. The Human Connectome Project resting state fMRI dataset [67] was used for this purpose (see Supporting Information). The raw data was pre-processed by applying a 3rd order Butterworth bandpass filter (0.01-0.08 Hz), then cutting 200 samples from the start and the end of the time series to remove potential filtering artefacts (leaving 800 samples for the analysis). In order to build a scenario where the null hypothesis held, 10 different random regions of interest (ROIs) were selected from different random subjects, such that the corresponding time series were expected to be independent of each other. The network inference was performed with the same settings used in the null test on synthetic data but employing the nearest-neighbour estimator, since the real data could not be assumed to follow a Gaussian distribution. The results on fMRI data are presented in Figure 8 and are consistent with the previous results on synthetic data (Figure 4). Unless appropriate measures are taken, the strong autocorrelation typically found in real data would result in an inflated false positive rate for short time series (an effect already observed by Barnett and Seth [68] when using Granger causality). The issue is addressed in IDTxl by means of the dynamic correlation exclusion, also known as Theiler window [69, 70], as originally suggested for transfer entropy estimation by Schreiber [7]. The idea is to exclude the closest points in time from the nearest-neighbour search which is necessary for the estimation of the transfer entropy (when using nearest-neighbour estimators). The autocorrelation decay time (i.e., the shortest time shift such that the autocorrelation function drops by a factor of 1/e with respect to the zero-shift value [13]) is used as a heuristic to adapt the size of the Theiler window to the data. FIG. 8. Validation of false positive rate for a single target (tFPR) on real fMRI data. The points indicate the average false positive rate over 50 repetitions of the experiment using random regions (N = 10) from different subjects. The horizontal marks indicate the corresponding 5th and 95th percentiles of the expected range. These were computed empirically from the distribution of the random variable (cid:104)Xj/N(cid:105), where Xj ∼ Binomial(N, αmax) are i.i.d. random variables, and the angular brackets indicate the finite average over 50 repetitions. The identity function is plotted as a reference (dashed line). C. Alternative visualisations of performance scaling by network and sample size Figure 9 replots the precision and recall from Figure 3 using different subplots for each sample size, and additionally shows the specificity. Similarly, Figure 10 replots the precision and recall from Figure 5 using different subplots for each sample size. 103102max103102FalsepositiverateN=10Identity 21 FIG. 9. Precision, recall, and specificity for different network sizes, sample sizes, and dynamics. Left: Vector autoregressive process; Right: Coupled logistic maps. Each row corresponds to a different time series length (Top: T = 100; middle: T = 1000, bottom: T = 10 000). The results for 10 simulations from different initial conditions are shown (low-opacity circles) in addition to the mean values (solid circles). 20406080100Networksize0.00.20.40.60.81.0T=100PrecisionRecallSpecificity20406080100Networksize0.00.20.40.60.81.0T=1000PrecisionRecallSpecificity20406080100Networksize0.00.20.40.60.81.0T=10000PrecisionRecallSpecificity20406080100Networksize0.00.20.40.60.81.0T=100PrecisionRecallSpecificity20406080100Networksize0.00.20.40.60.81.0T=1000PrecisionRecallSpecificity20406080100Networksize0.00.20.40.60.81.0T=10000PrecisionRecallSpecificity 22 FIG. 10. Precision-recall trade-off for different statistical significance levels. The plots show the results for different dynamics (Left: Vector autoregressive process; Right: Coupled logistic maps), different time series lengths (top to bottom: T = 100, 1000, 10 000), and different network sizes (N = 10, 40, 70, 100). The error bars indicate the standard deviation over 10 simulations from different initial conditions. max0.00.20.40.60.81.0PrecisionT=100N=10N=40N=70N=100maxRecallT=100N=10N=40N=70N=100max0.00.20.40.60.81.0PrecisionT=1000N=10N=40N=70N=100maxRecallT=1000N=10N=40N=70N=100103102101max0.00.20.40.60.81.0PrecisionT=10000N=10N=40N=70N=100103102101maxRecallT=10000N=10N=40N=70N=100max0.00.20.40.60.81.0PrecisionT=100N=10N=40N=70N=100maxRecallT=100N=10N=40N=70N=100max0.00.20.40.60.81.0PrecisionT=1000N=10N=40N=70N=100maxRecallT=1000N=10N=40N=70N=100103102max0.00.20.40.60.81.0PrecisionT=10000N=10N=40N=70N=100103102maxRecallT=10000N=10N=40N=70N=100
1801.04631
2
1801
2019-09-23T21:55:14
Noise Sharing and Mexican Hat Coupling in a Stochastic Neural Field
[ "q-bio.NC" ]
A diffusion-type coupling operator biologically significant in neuroscience is a difference of Gaussian functions (Mexican Hat operator) used as a spatial-convolution kernel. We are interested in pattern formation by \emph{stochastic} neural field equations, a class of space-time stochastic differential-integral equations using the Mexican Hat kernel. We explore, quantitatively, how the parameters that control the shape of the coupling kernel, coupling strength, and aspects of spatially-smoothed space-time noise, influence the pattern in the resulting evolving random field. We confirm that a spatial pattern that is damped in time in a deterministic system may be sustained and amplified by stochasticity. We find that spatially-smoothed noise alone causes pattern formation even without direct spatial coupling. Our analysis of the interaction between coupling and noise sharing allows us to determine parameter combinations that are optimal for the formation of spatial pattern.
q-bio.NC
q-bio
Noise Sharing and Mexican Hat Coupling in a Stochastic Neural Field 1Department of Mathematics, University of Southern California, Los Angeles, CA, USA Peter H. Baxendale1 2Department of Mathematics, University of British Columbia, Vancouver, BC, Canada Priscilla E. Greenwood2 3Department of Psychology and Brain Research Centre, 2136 West Mall, University of British Columbia, Vancouver, BC, V6T 1Z4 Canada∗ Lawrence M. Ward3 (Dated: September 25, 2019) A diffusion-type coupling operator biologically significant in neuroscience is a difference of Gaussian functions (Mexican Hat operator) used as a spatial-convolution kernel. We are interested in pattern formation by stochastic neural field equations, a class of space-time stochastic differential-integral equations using the Mexican Hat kernel. We explore, quantitatively, how the parameters that control the shape of the coupling kernel, coupling strength, and aspects of spatially-smoothed space-time noise, influence the pattern in the resulting evolving random field. We confirm that a spatial pattern that is damped in time in a deterministic system may be sustained and amplified by stochasticity. We find that spatially-smoothed noise alone causes pattern formation even without direct spatial coupling. Our analysis of the interaction between coupling and noise sharing allows us to determine parameter combinations that are optimal for the formation of spatial pattern. Keywords: neural field equation, Mexican Hat coupling, stochastic process, spatial pattern, spatially-smoothed noise. PACS: 89.75.-k,, 89.75.Da, 05.45.Xt, 87.18.-h MSC: 34C15, 35Q70 I. INTRODUCTION In this paper we explore the formation of pattern by a stochastic neural field equation with simple damping as its reaction term and with both i.i.d. and shared noise. The existence and possible sources of shared noise cor- relations in stochastic neural models have been studied recently by Doiron et al. [1], and by Meyer, Ladenbauer, and Obermayer [2]. We, on the other hand focus on the effect of shared or correlated noise on pattern formation. We have been shown by Hutt and colleagues [3] that spatial pattern, embedded in deterministic space-time dynamics but immediately damped, may be excited by noise. Butler and Goldenfeld [4, 5], and McKane, Bian- calani and Rogers [6] showed the existence of excitable spatial modes that, when noise was added, were revealed in power spectral densities. The knowledge of this noise- facilitated source of pattern, also observed in biological systems, motivated us to explore how certain sample path properties of evolving stochastic neural fields depend on the parameters in a basic example. We look at parame- ters that control strength of coupling or local interaction in the field, and on the extent of local sharing, or smooth- ing, of noise. The sample path properties we look at are: how pat- tern grows with coupling strength, how pattern is re- vealed and sustained by noise when field interaction has no excitable modes, how coupling interacts with noise smoothing, giving rise to distinct patterns for optimum smoothing, and then yielding distorted pattern with "too much smoothing." A striking result is that spatial smoothing of noise, alone, without direct neural field in- teraction, produces pattern. Simulation of the time evolution of one-dimensional fields yields insight about typical sample path behaviour of the evolving random field and illustrates the analytical conclusions we present. We use two measures of spatial pattern. One is the spatial FFT amplitude as a stochas- tic process in time (note, for comparison to other stud- ies, that power spectral density is the square of the FFT amplitude; here we compute spatial FFT amplitudes di- rectly and from simulations). The second is a function of space, which we call F , that allows direct observation of dominant frequency, again a stochastic process in time. We believe that display of sample paths of both of these processes is innovative here, and that they will prove to be valuable in future studies of the model we study and of other stochastic neural fields. What has come to be called a neural field equation is an integro-differential equation of the form dY (t, x) = −Y (t, x)+ cw(x−y)S(Y (t, y))dy dt, (1) (cid:20) (cid:90) R (cid:21) ∗ Corresponding author. Email: [email protected], Tel.: +1 604 822 6309, Fax: +1 604 822 6923. where Y is an R1-valued state variable for a neural sys- tem, w is a coupling operator, for example the Mexican Hat convolution kernel, c, is a constant called the cou- (cid:20) (cid:90) L 0 (cid:21) dt (2) pling strength, and S is a scaling functional, typically a sigmoid, which keeps Y bounded in the diffusion term. In what follows we take S to be the identity within a certain bounding range. We study a stochastic version of (1), wrapped with length L, dY (t, x) = − Y (t, x) + cw(x − y)Y (t, y)dy + σdG(t, x), where σ is a constant diffusion coefficient, c ≥ 0 is a con- stant, and G(t, x) is a Gaussian process that may depend on both time, t, and space, x. We study the interaction of coupling and noise smoothing in (2) theoretically in continuous space and time, and then discretize both for simulations. Recently, the study of such equations, with continu- ous state variable, has been put on a rigorous footing by Faugeras and Inglis [7]. Conveniently for the study of neural field equations that generate spatial pattern, they singled out the difference-of-Gaussians coupling op- erator (often called the Mexican Hat operator) as one that satisfies established conditions for the existence and uniqueness of a solution. As will be seen, this coupling operator generates pattern when used in equations like (2). It is sometimes preferred to the Laplacian as a cou- pling operator [8]. In what follows we compute conditions for the inter- action of the dominant modes produced by coupling and noise smoothing in the stochastic model (2). When no excitable modes are present the spatial pattern gener- ated by the dominant modes is damped, but added noise reveals and sustains the damped spatial pattern. Noise smoothing can either render the revealed pattern more clearly, or distort it, depending on the extent of the smoothing. New here is an analytic and graphical study of so- lutions of (2) with smoothed noise. We are interested in spatial pattern revealed by noise in stochastic sam- ple paths from the evolving random field of the neural field equation as parameters are varied. We explore the relation between the coupling strength constant, c, and a parameter of noise smoothing, η, the standard devia- tion (SD) of a Gaussian smoothing kernel. We consider separately the ranges of c where the eigenvalues are all negative and where the maximum eigenvalue is positive. In our computations a specific value of c separates these ranges. Also new is the result that smoothed noise, by itself with c = 0 in (2), can produce spatial patterns sim- ilar to those arising from excitable spatial modes created by the Mexican Hat coupling. II. INPUT NOISE In modeling a spatially discrete stochastic neural field, a default choice (e.g., [9]) has been to introduce an inde- pendent Brownian component for each x. If the locations 2 are tightly packed, however, as in neural tissue, the same input noise may be shared in a neighborhood of loca- tions. This is because the nearer neurons are to each other in the brain the more similar is the set of synaptic inputs they receive. Neurons communicate more exten- sively with each other the closer they are to each other because the number and strength of synaptic connections between neighboring neurons tends to decline with the distance between them [10, 11]. Given that shared noise would be mostly synaptic noise [11, 12], the amount of shared noise between two neurons should decline with the amount of effective connectivity between them. We are interested in the effect on spatial patterns of the size of the neighborhood in which some of the same input noise is felt. In our exploration we have included cases where noise is independent at each location, x, and cases where noise is averaged over neighborhoods of various sizes. We refer to this as spatial smoothing of noise. It has been shown that such noise allows for (Holder) continuous so- lutions to a broad class of stochastic integro-differential equations, including equations such as (2) [7, 13 -- 15]. In simulations, Section IV, for spatial smoothing we used a Gaussian kernel. The Gaussian smoothing kernel was convolved with i.i.d. Gaussian noises at each iteration of the evolving spatial field. III. INTERACTION OF COUPLING AND NOISE SMOOTHING IN THE NEURAL FIELD EQUATION We capture the interaction of coupling and noise smoothing in terms of the Fourier component processes associated with (2). We begin with (2), taking w and r to be general functions. Later we specialize to Mexican Hat and Gaussian functions. The noise G(x, t) is Gaussian with mean 0 and covariance E [G(t, x)G(s, y)] = min(s, t)r(x − y). (3) Here r is a general positive semi-definite function. We denote the Fourier transforms of the coupling kernel w and the noise smoothing function r by and W (k) = R(k) = e−ikxw(x) dx, k ∈ R, e−ikxr(x) dx, k ∈ R. (cid:90) ∞ (cid:90) ∞ −∞ −∞ riodic. That is, replace w(x) by w(x) =(cid:80) and r(x) =(cid:80) The spatial variable x is in an interval [0, L] with pe- riodic boundary conditions, equivalently on the circle R/LZ, so long as the functions w and r are made pe- n∈Z w(x + nL) n∈Z r(x + nL) and drop the 'hat' notation. In fact the widths of the support of r and w will be less than L. (cid:90) L (cid:90) L 0 0 The periodic function w(x) has Fourier coefficients W (2πk/L) = e−2πikx/Lw(x) dx, k ∈ Z, (4) and the periodic function r(x) has Fourier coefficients R(2πk/L) = e−2πikx/Lr(x) dx, k ∈ Z. (5) Moreover r is positive semi-definite. We consider the model (2), where the noise G(x, t) is given by (3). A Fourier series expansion of the solution Y (t, x) of (2) allows us to write it in terms of a family of Ornstein- Uhlenbeck processes indexed by spatial frequencies (wave numbers, k). We can then compute the expected squared amplitudes of these O-U processes. There are standard 2- dimensional Brownian motions {Ck(t) : t ≥ 0} for k ≥ 1 and a standard scalar Brownian motion {C0(t) : t ≥ 0}, all mutually independent, such that the solution Y (t, x) of (2) is given by Proposition (cf. [17]). (cid:16) ak(t)e2πikx/L(cid:17) (cid:112)R(0) dC0(t) (7) (6) Y (t, x) = a0(t) + 2Re where the scalar process a0(t) satisfies σ√ da0(t) = [−1 + cW (0)] a0(t) dt + L and for k ≥ 1 the complex processes ak(t) satisfy dak(t) = [−1 + cW (2πk/L)] ak(t) dt + (cid:112)R(2πk/L) dCk(t). σ√ 2L Proof. Substituting the Fourier series expansion Y (t, x) = ak(t)e2πikx/L (cid:88) k≥1 (cid:88) k∈Z into (2) gives(cid:88) −(cid:88) (cid:90) L k∈Z k∈Z dak(t)e2πikx/L = ak(t)e2πikx/Ldt (cid:88) k∈Z w(x − y) + c 0 + σdG(t, x). Since (cid:90) L ak(t)e2πiky/Ldydt w(x − y)e2πiky/L dy (cid:90) L 0 = e2πikx/L 0 = e2πikx/LW (2πk/L) w(x − y)e2πik(y−x)/L dy 3 we get(cid:88) (cid:88) k∈Z = k∈Z dak(t)e2πikx/L (cid:2)−1 + cW (2πk/L)(cid:3)ak(t)e2πikx/L dt + σdG(t, x). Now multiply by e−2πi(cid:96)x/L and integrate with respect to x. We get Lda(cid:96)(t) = L [−1 + cW (2π(cid:96)/L)] a(cid:96)(t) dt + σdB(cid:96)(t), equivalently dak(t) = [−1 + cW (2πk/L)] ak(t) dt + σ L dBk(t) (9) where Bk(t) = (cid:90) L 0 e−2πikx/LG(t, x) dx, k ∈ Z. (10) Assuming that the initial condition Y (0, x) is real, we have a−k(0) = ak(0). Since B−k(t) = Bk(t) for all t, we have a−k(t) = ak(t) for all t ≥ 0. Therefore and it suffices to study ak(t) for k ≥ 0. ak(t)e2πikx/L + a−k(t)e−2πikx/L = 2Re(cid:0)ak(t)e2πikx/L(cid:1), may write B0(t) = (cid:112)LR(0)C0(t) and Bk(t) = (cid:112)(L/2)R(2πk/L)Ck(t) for k ≥ 1, where the processes Ck are as stated in the proposition. For k ≥ 1 the equation (9) can then be rewritten as (7) and (8). in Appendix A, we calculations Using the (8) Corollary 1. Suppose Y (0, x) = A0 + 2 Ak cos(cid:0)2πkx/L + φk (cid:1) (cid:88) k≥1 Let a0(t) be the solution of (7) with initial condition a0(0) = A0, and for k ≥ 1 let ak(t) be the solution of (8) with initial condition ak(0) = Akeiφk . Write ak(t) = Ak(t)eiφk(t). Then Ak(t) cos(cid:0)2πkx/L + φk(t)(cid:1). Y (t, x) = a0(t) + 2 (cid:88) k≥1 Thus the period k part of Y (t,·) has amplitude 2Ak(t) and phase φk(t) determined by the Ornstein-Uhlenbeck process ak. Look more closely at the process ak(t). We consider k ≥ 1 (the k = 0 case is similar). For ease of notation write (8) as dak(t) = λkak(t) dt + σk dCk(t). This complex-valued SDE has solution ak(t) = eλktak(0) + σk ≡ eλktak(0) + Nk(t), eλk(t−s)dCk(s) (11) (cid:90) t 0 [16]. Here and λk = −1 + cW (2πk/L) (cid:112)R(2πk/L) σk = σ√ 2L for k ≥ 1 (and slight modifications to the formulas if k = 0). We can now outline our overall strategy. Because the integral (2) is over a bounded set, the eigenvalues of the operator in (2) are separated, and the mode k of the dominant eigenvalue determines the spatial frequency of the spatial pattern. For each pair of parameters (c, η), which will determine the coupling strength and the width of the noise smoothing, defined in section III B, we are able to evaluate a functional of the distribution of the process Yc,η that is maximal at that value of (c, η) for which kc,η is the dominant mode. This functional, the expected squared amplitude of the kth mode of Yc,η, can be computed in terms of the drift and diffusion coeffi- cients, λk, σk, of the process ak(t) defined by (7), (8). Since Ck is standard 2-dimensional Brownian motion, then Nk(t) is 2-dimensional Gaussian with mean zero and covariance matrix e2λk(t−s)ds I2 = k(e2λkt − 1) σ2 2λk I2 ≡ vk(t)I2 (cid:18)(cid:90) t 0 σ2 k (cid:19) where I2 denotes the 2 × 2 identity matrix. Write ak(0) = αk + iβk. Then the real part of ak is normal with mean eλktαk and variance vk(t), and the imaginary part of ak(t) is normal with mean eλktβk and variance vk(t), and the real and imaginary parts are independent. Then [16] EAk(t)2 = Eak(t)2 = e2λkt(α2 k + β2 k) + 2v2 k = e2λktAk(0)2 + k(e2λkt − 1) σ2 λk . (12) This formula is valid for finite t regardless of the sign of λk. Suppose that λk > 0 (that is, cW (2πk/L) > 1) for some k. Then ak(t) will exhibit exponential growth. We treat seperately the cases where all λk < 0 and those where there exist k such that λk > 0. A. Relationship between σ2 k and λk Suppose that cW (2πk/L) < 1 for all k, that is all λk < 0. Then the effect of the initial condition dies away, and each Nk(t) converges to a stationary Ornstein- Uhlenbeck process. The real and complex (independent) components of Nk(t) each satisfy a scalar equation dZt = λkZtdt + σkdWt which has stationary distribution N (0, σ2 that, from (12), k/(−2λk)), so 4 (13) . EAk(t)2 ∼ ENk(t)2 = = = σ2R(2πk/L) σ2 k−λk 2L(1 − cW (2πk/L)) σ2 2L (1 − cW (2πk/L)) R(2πk/L) When do we see the kth mode as dominant in a sta- tionary solution? We should look for a parameter region where cW (2πk/L) < 1 and thus λk = −1+cW (2πk/L) < (1 − cW (2πk/L)) 0 for all k and also the max value of R(2πk/L) is noticeably larger than its other values. Notice that when λk = 0 there exists a critical point at which center manifold theory applies [17, 18]. In this paper we do not address center manifold theory. We first confine ourselves to the study of the case λk < 0 for all k, where we have an explicit expression for the relationship, (13), between the noise smoothing and the Mexican Hat influences on the solutions to (2). We then find that, in spite of the increasing exponential in (12) when some λk > 0, we can still use the ratio in (13) to predict the optimal pairings of the width of the noise smoothing, η, and the coupling strength, c, to obtain a clear but transient pattern in some range of t. B. Gaussian Noise Smoother As stated earlier, our noise G(t, x) is Gaussian with mean 0 and covariance E [G(t, x)G(s, y)] = min(s, t)r(x − y). If Z(t, x) is a Gaussian family, Brownian in time and uncorrelated in space, then E[Z(t, x)Z(s, y)] = min(s, t)δ(x − y), and, if g is symmetric, then the smoothed noise (cid:90) R Z(t, y)g(x − y) dy G(t, x) = has E[G(t, x)G(s, y)] = min(s, t)r(x − y) where r(x − y) = (g ∗ g)(x − y). Here we used the sym- metry of g. In particular if g is the density of N (0, η2) then r is the density of N (0, 2η2), so that exp(−x2/4η2). r(x) = √ 1 2η π The Fourier transform of r(x) is thus R(k) = exp(−η2k2). (14) √ 2√ π In Section V we explore by simulation the effect on the emerging spatial pattern of changing the standard deviation, η, of the Gaussian noise smoother. We note that in addition to the effect of the noise smoother on the dominant Fourier mode, (13), there is also a variance reduction effect that is dependent on η as well. That is, the variance, σ2 s , of the smoothed noise for each process, ak, is approximated by (cid:90) +∞ (cid:20) exp(−x2/2η2) (cid:21)2 √ η s ≈ σ2 σ2 σ2 √ σ2 2π −∞ dx = √ (15) s < σ2. This implies that when η > 1/(2 We should thus expect that smoothed noise might be more effective at revealing spatial patterns than would i.i.d. noise alone, at least when η > 0.28 (c.f., [19]). π) = 0.28, σ2 3.54η 2η = π . C. Mexican Hat Coupling Operator We chose a form for w(x), the difference of the two Gaussian functions, that expresses the common biological observation that there is excitation within a small neigh- bourhood around each location, and inhibition in a some- what larger neighbourhood around the excitation. An- other way to achieve this effect is to multiply by a Gaus- sian function and then operate with a Laplacian [20]; yet another alternative operator that involves a squared Laplacian is used in [8]. The existence and identity of excitable spatial modes that lead to spatial pattern in neural fields has been studied also using other approaches (see, e.g., [3] and for a review see [21]). To be explicit, the Mexican Hat operator [22] is defined as w(x) = b1 exp(−(x/d1)2) − b2 exp(−(x/d2)2) and its Fourier transform is √ W (k) = π b1d1 exp − (d1k)2 4 (cid:20) (cid:21) −b2d2 exp (cid:20) − (d2k)2 4 (16) (cid:21)(cid:35) . (cid:34) (cid:32) (17) So that w(x) has indeed a Mexican Hat shape we need b1/b2 > 1 and d2/d1 > 1. Under the additional condition that b2d3 1, then kmax, the wave number for which W (k) is largest, is given by 2 > b1d3 kmax = 4(d2 2 − d2 1)−1 ln . (18) (cid:19)3(cid:35)(cid:33)0.5 (cid:34) (cid:18) d2 d1 b2 b1 5 use the Mexican hat operator with b2 = d1 = 1, b1 = 1.1, d2 = 1.2. With these values kmax = 2.026 and W (kmax) = 0.2134 so that −1 + cW (kmax) < 0 for all c < 4.685. Note that we are treating η as a parameter of (14) so that it can indeed take the value of 0. It can be seen in Figure 1 that when η = 0 the value R(k) (1 − cW (k)) of c. It follows that the dominant mode k, that is the is maximized at kmax = 2.026 for all values R(2πk/L) (1 − cW (2πk/L)) integer value of k which maximizes , is approximately kmaxL/(2π) = 0.322L ≈ 8 (for the value of L used in our simulations). As the width η of the spa- tial noise smoother increases, the dominant mode tends to occur at lower values of k, indicating fewer spatial cy- cles on the ring. The decline of the dominant mode with increasing noise smoothing width occurs for lower values of η when we consider lower values of c. For the largest values of η the dominant mode is at k = 0, as mentioned earlier. IV. SIMULATIONS An approach consistent with the preceding theory would be to simulate each of the Fourier coefficients ak(t), given by (11), using the calculations above. For compu- tational purposes we would simulate only a finite number of them, say 0 ≤ k ≤ K where K is chosen so that EA2 given by (13) is small for all k > K. k In fact we chose to simulate the space-time neural field equation (2) more directly and implemented a discrete (both space and time) approximation using the Euler- Maruyama procedure for obtaining a numerical solution of stochastic difference equations. This approach has been shown to converge rapidly to a close approxima- tion of the continuous solution [23]. For a collection of equally spaced points {xj} write Y (t, xj) = Yj(t) and G(t, xj) = Gj(t). The Euler-Maruyama method gives: Yj(t + ∆t) − Yj(t) =(cid:2) − Yj(t) + ch (cid:88) wj(cid:96)Y(cid:96)(t)(cid:3)∆t + σ Gj(t + ∆t) − Gj(t) (cid:16) (cid:17) (cid:96) (19) where wj(cid:96) represents the discretized Mexican Hat oper- ator and h denotes the spacing between points xj. The factor h is introduced so that the sum in (19) will be a good approximation to the integral in (2). D. Interaction of k, c, η Because of its shape-determining role in (13) we pro- duce graphs, Figure 1, displaying the way in which the expression R(k) (1 − cW (k)) varies with k, c and η. We We used n = 128 points spaced at distance h = 0.2 apart in a 1D spatial array, and implemented a periodic boundary so that the values of xj, and the corresponding components Yj(t) of (19), can be thought of as forming a discrete ring of 128 components. This corresponds to a A. Simulation implementation 6 FIG. 1. Effects of spatial noise smoother standard deviation, η, on the ratio R(k)/(1 − cW (k)) for various values of c. Note that the largest values of the ratio occur in the top curve in each plot, where η = 0, with curves descending in value with increase in η. Vertical line is at k = 2, near the maximum value for all values of c when η = 0. wrapping length L = nh = 128 × 0.2 = 25.6 in equation (2). The 128 components were coupled by the Mexican Hat operator (16) with parameter values b2 = d1 = 1, b1 = 1.1, d2 = 1.2. With these values the function w(x) had effective width from -3 to 3, being very near zero out- side this interval, so that our discretized Mexican Hat op- erator wj(cid:96) was effectively 31 spatial locations wide. That is, the choice h = 0.2 was convenient for the implemen- tation of the Mexican Hat so that it operated only on 31 of the 128 components around each position in the wrapped 1D spatial array. We chose the width of 31 for the Mexican Hat operator because (a) each single Mexi- can Hat-type neural coupling in a neural system should span only a fraction of the size of the system, and (b) 31 spatial locations, while being only a fraction of the sys- tem size of 128, is large enough to model a Mexican Hat operator that couples several excitatory and inhibitory components. A "biologically" based choice would depend on the relative size of an observed neural field to the span of a neighborhood of a neural location that includes both (a) (e) (c) (d) (b) (f) excitatory and inhibitory neighbors. We implemented space-time noise as described in Sec- tion II. In the spatially i.i.d. case the noise processes Gj(t) were independent standard scalar Brownian mo- tions, while for smoothed noise with parameter η the dis- tribution of Gj(t) = G(t, xj) was as described in Section III B. That is, on each iteration independent samples of Gaussian noise for each component were combined for each component by a normalized Gaussian kernel, with standard deviation η, centered at that component. In our simulations, presented in Section V, we studied spa- tially i.i.d. noise, and values for η = 0.15, 0.5, 0.67, and 1.3, corresponding roughly to kernels effectively about 1 (i.e., i.i.d., no smoothing), 5, 15, 21, and 39 spatial lo- cations wide. These widths span a range from spatially i.i.d. to a smoothing that combines noise from locations over nearly 1/3 of the total ring. We solved the stochastic difference equation (19), also with σ = 0, iteratively and typically for 10,000 time steps with ∆t = 0.00005, corresponding to a time interval of length 0.5. In a few cases, to be indicated, we extended the simulation to 500,000 iterations and t = 25, and in a few others we used ∆t = 0.0025 in order to reach t = 25 in only 10,000 iterations. The initial values Yj(0) for each component were independent random variables cho- sen uniformly in the interval [0.5, 0.501]. The initial per- turbation from the constant 0.5 is necessary to observe pattern in the evolving coupled field in the absence of noise [24]. For cases where all λk < 0 we calculate the amplitude of the dominant harmonic and compare that with the am- plitude of this harmonic in simulations. For cases where the maximum λk > 0 we illustrate spatial pattern for each set of parameter values for a representative realiza- tion of the paths of all 128 components of (19) with the ring flattened out. We also display for those realizations the Fourier amplitudes (from a Fast Fourier Transform, FFT) of the spatial frequencies as a stochastic process in t, and a second measure we term F . For both the FFT amplitudes and the computation of F we coarse- grained time, considering 500-iteration time blocks: 1- 500, 751-1250, 1751-2250, ...,7751-8250, 8751-9250, 9501- 10000. Note that the gaps between the first two blocks and between the final two blocks are 250 and the gaps between the remaining contiguous timeblocks are all 500. For the FFT amplitudes we averaged Y(t,x) for each com- ponent over each 500-iteration block and then computed the FFT on the resulting spatial array. F is a function of the time-block parameter τ and (cid:96), written as F (τ, (cid:96)) = 1 500 1 m s ∈ timeblock τ j=1 Yj+(cid:96)(s) − Yj(s), (20) where (cid:96) is a spatial offset and m is the distance across the array for which we are computing F . In our computations m was fixed at m = 64 because the period of the spatial pattern never exceeded this value. In the computation (cid:96) is increased progressively across the spatial array. Thus, (cid:88) m(cid:88) 7 whenever the difference Yj+l(s)−Yj(s) is large, the value of F is correspondingly increased. Local maxima in the plot of F (τ, (cid:96)) occur wherever the spatial offset (cid:96) matches half the period of a spatial pattern. The presence of clear maxima in F indicates the presence of a periodic spatial pattern, and the form of the pattern in F displays the pattern in the Yj(t) but sometimes more clearly because of the averaging implicit in the computation of F . V. RESULTS A. All λk < 0 In this section we compare representative calculations based on the theory outlined earlier with simulations of (19) for cases where all λk < 0, and one case where λk > 0 for illustrative purposes. In our numerical work, the value of c that separates the case where all λk < 0 and where λk > 0 for some k is c ≈ 4.685. 1. No noise Before we launch into our study of the joint effects of smoothed noise and the Mexican Hat on the solutions to (2), we compute the behaviour of (2) without noise. The values for the Yj(0) are i.i.d. uniform in [0.5, 0.501]. √ Thus each Yj(0) has mean value µ = 0.5005 and standard deviation γ = 0.001/ 12 ≈ 0.0003. Since n−1(cid:88) Yj(0) = Y (0, jL/n) = ak(0)e2πik(jL/n)/L n−1(cid:88) = k=0 we obtain k=0 ak(0)e2πijk/n n−1(cid:88) j=0 1 n ak(0) = Yj(0)e−2πijk/n. 128 = 0.000026. √ 12)/ √ n = (0.001/ Taking k = 0, we see that a0(0) is real with mean µ and variance γ2/n. For 1 ≤ k ≤ n − 1 we have E[ak(0)] = 0 and Eak(0)2 = γ2/n. Thus typically Ak = ak(0) is of √ order γ/ For c < 4.685, −1 + cW (2πk/L) < 0 for all k and so no persistent pattern is predicted. Consider, for ex- ample, c = 4.5. The eigenvalue −1 + 4.5W (2πk/L) is maximized when 2πk/L ≈ kmax = 2.0263, that is for k = 8. Thus the eigenvalue for the 8th harmonic is −1 + 4.5W (16π/L) ≈ −1 + 4.5W (kmax) = −1 + 4.5 × 0.22 = −1 + 0.99 = −0.01. Over the time interval [0, 0.5] the 8th harmonic "grows" (actually damps) by a factor of e(−1+4.5W (16π/L))×0.5 ≈ e−0.005 = 0.995. Given the initial amplitude of the 8th harmonic is ap- proximately 0.000026, the typical final amplitude of the 8th harmonic would be 0.995 × 0.000026 = 0.00002587. We simulated (19) with c = 4.5 and the average FFT am- plitude of the 8th harmonic over the final 500 iterations (up to t=0.5) from 10 realizations (different realizations of the starting values) was 0.0000227, very close to the predicted value for the dominant 8th harmonic calculated from (12) with no noise and c = 4.5. To illustrate calculations when λk > 0 for some k we now consider a value of c = 15. In this case the eigenvalue for the 8th harmonic is −1 + 15W (16π/L) ≈ −1 + 15W (kmax) = −1 + 15 × 0.22 = −1 + 3.3 = 2.3. Over the time interval [0, 0.5] the 8th harmonic in this case does grow by a factor e(−1+15W (16π/L))×0.5 ≈ e1.15 = 3.16. Since the typical initial amplitude of the 8th harmonic is approximately 0.000026, then the typical final amplitude of the 8th harmonic will be 3.16 × 0.000026 = 0.000082. We simulated (19) with c = 15 and Figure 2a displays the average amplitudes of the spatial harmonics over 10 realizations (different realizations of the starting values). The average value of the FFT amplitude at a spatial fre- quency of 8 for these 10 realizations of (19) is 0.000082, exactly the predicted value for the dominant 8th har- monic calculated from (12) with no noise and c = 15. 2. I.I.D. noise For i.i.d. noise the discrete form of the Fourier ex- pansion is valid with r(x − y) replaced by rjk = δjk. This in turn gives a version of the Proposition (valid for x = xj = jL/n) with R(2πk/L)/L replaced by 1/n. As an application of the theory we can use (12) with random initial values: k(e2λkt − 1) σ2 λk E[Ak(t)2] = e2λktE[Ak(0)2] + k = σ2/(2n) for k ≥ 1. With c = 4.5 we have where σ2 k/(−λk) λk < 0 for all k ≥ 1, and then E[Ak(t)2] → σ2 k/(−λk) is as t → ∞. With σ = 1 the limiting value σ2 maximized when k = 8 and has max value 0.092. This corresponds to A8(t) of order 0.092 = 0.30, for large t. But this is the limiting value. The value at time t = 0.5 is much smaller. The first term contributing to E[Ak(t)2] is at most E[Ak(0)2] = (0.000026)2. The second term is k(eλk − 1) σ2 which is maximized at k = 8 with max value 0.0038. Together E[Ak(0.5)2] ≤ 0.0039 for all k ≥ 1. Thus Ak(0.5) is of order at most 0.0039 ≈ 0.06. √ λk √ 8 FIG. 2. FFT amplitudes averaged over 10 realizations with ∆t = 0.00005. (a) No noise, c = 15, σ = 0, t = 0.5. (b) I.I.D. noise, c = 4.5, σ = 1.0, t = 0.5. (c) I.I.D. noise, c = 4.5, σ = 1.0, t = 25. (d) Smoothed noise, c = 4.5, σ = 1.0, η = 0.5, t = 0.5. (e) Smoothed noise, c = 4.5, σ = 1.0, η = 0.5, t = 25. In all cases the FFT amplitudes are computed on the average Yj(t) over the final 500 iterations (up to t = 0.5 or t = 25), kmax = 2.0263, and the number of cycles in L/2π (spatial frequency in graphs) is 8, as described in the text. We simulated (19) with c = 4.5 and Figure 2b displays the average FFT amplitudes for 10 realizations (different realizations of the starting values and the noise). The av- erage value of the FFT amplitude at a spatial frequency of 8 and time period 11 (average over final 500 itera- tions up to t = 0.5) for these 10 realizations of (19) is 0.059, very close to the predicted value for the dominant 8th harmonic from the theory. Notice, however, that the dominant frequency doesn't stand out very well from the noise at other spatial frequencies in Figure 2b. Thus i.i.d. noise doesn't substantially enhance the weak pattern ex- pected from c = 4.5 at this short time interval. To see the effect of i.i.d. noise over a longer time interval we simulated 10 realizations of (19) with c = 4.5, ∆t = 0.00005 over a time interval from t = 0 to t = 25 (500,000 iterations). Figure 2c shows the aver- age FFT amplitude over the final 500 iterations. The average FFT amplitude at a spatial frequency of 8 is ap- proximately 0.27, close to the predicted value of 0.30 for large t from the theory. In order to observe the stochastic E"C"B"A"D"(a) (e) (c) (d) (b) paths of Yj(t), F and FFT amplitude, we also simulated a single realization of (19) with c = 4.5 over a time in- terval from t = 0 to t = 25 but with ∆t = 0.0025. Figure 3a shows the results of this simulation. After the longer time interval the pattern more clearly stands out from the noise, consistent with the FFT amplitude shown in Figure 2c. There is still considerable noise, however, ev- ident in the spatial pattern of Yj(t). 3. Coupling and noise smoothing The interaction between smoothing width and cou- pling strength for λk < 0 in regard to dominant modes is described by (13). As depicted in Figure 1, the ra- tio R(k) (1 − cW (k)) predicts the dominant Fourier mode for various values of η and c. Noticeably, for any value of c, as η increases the dominant mode eventually goes to 0. Also, noticeably, for higher values of c this approach to 0 occurs at higher values of η. In addition, the variance reduction property of the spatial noise smoothing, (15), operates for values of η > 0.28, whereas for η < 0.28 variance is actually increased by the 'smoothing.' Thus, the noise smoothing has two somewhat conflicting effects: decreasing the variance of the noise-revealed spatial pat- tern induced by the Mexican Hat operator, but also tend- ing to distort it towards lower spatial frequencies. And this tradeoff also depends on the value of the coupling strength; higher values of c allow the variance reduction effect to override the dominant mode reduction. To calculate an example of the application of the the- ory for smoothed noise we can use (12) with random ini- tial values as in the case of i.i.d. noise. With c = 4.5.η = k/(−λk) has max value 0.14. This 0.5, the limiting value σ2 corresponds to A8(t) of order 0.14 = 0.37, for large t. √ But again this is the limiting value. The value at time t = 0.5 is much smaller. The first term contribut- ing to E[Ak(0.5)2] is at most E[Ak(0)2] = (0.000082)2. The second term is . For η = 0.5 this is k(eλk − 1) σ2 λk maximized at k = 5 with max value 0.0083. Together E[Ak(0.5)2] ≤ 0.0084 for all k ≥ 1. Thus Ak(0.5) is of order at most 0.0084 ≈ 0.09. √ We simulated (19) with c = 4.5, σ = 1.0, η = 0.5 and Figure 2d displays the average spatial FFT amplitude re- sults of 10 realizations (different realizations of the start- ing values and of the noise). The average value of the FFT amplitude at a spatial frequency of 8 and time pe- riod 11 (average over final 500 iterations up to t = 0.5) for these 10 reactions of (19) is about 0.06, in line with the predicted value from the theory. Here, however, the dominant mode from the coupling stands out better from the surrounding spatial frequencies, except for the lower frequencies where the dominant mode of the smoother re- sides. The effect of the smoother is to suppress the noise at frequencies higher than that of the dominant mode, rather than to enhance the dominant mode, whose ampli- 9 tude is about that with i.i.d. noise. Note, however, that both with i.i.d. noise and smoothed noise, the amplitude of the dominant mode is substantially greater than that with no noise, even for much larger values of c in the no noise case. Thus, the noise amplifies the spatial pattern, and smoothed noise makes the pattern stand out from the noise. Because for short simulation times and λk < 0 the am- plitude processes of the Fourier coefficients and Y (x, t) will be small, we also simulated 10 realizations of the case c = 4.5, η = 0.5, σ = 1.0 over the time interval t = 25 with the same ∆t = 0.00005. For t = 25 the limiting value of the Fourier coefficient of the 8th harmonic for large t is approximately 0.37. Figure 2e displays the av- erage FFT amplitude from the 10 realizations, which is about 0.24, again in line with the theory. In addition, we simulated the case c = 4.5, η = 0.5, σ = 1.0 over the time interval from t = 0 to t = 25 but with ∆t = 0.0025 so that we could view the sample paths of Y (t, x), F and FFT amplitude. Figure 3b displays the results of this simulation. The spatial pattern is clearly evident over this longer time interval, again as predicted by the theory. Note that the pattern is quite smooth on the final iteration, in contrast to the pattern with i.i.d. noise. Again, the spatial smoothing suppresses the noise at higher frequencies, acting as a bandpass filter, rather than enhancing the spatial pattern itself. B. Max λk > 0 In this section we display the results of simulations of (19) with values of c such that λk > 0 for some k. Here (12) applies, so we expect exponential increases with time in the amplitudes of the spatial modes, and thus in the spatial pattern. Nonetheless, the interaction of the noise smoothing parameter and the coupling strength are accurately predicted by (13) and Figure 1. 1. No noise Figure 4 displays solutions of (2) with σ = 0, i.e., solu- tions to the deterministic version of (2) for the 128 spa- tial replicants. With the parameters b2 = d1 = 1, b1 = 1.1, d2 = 1.2, as before, we have W (0) ≈ −0.177 and kmax ≈ 2 and W (kmax) ≈ 0.22. For k = 0 we have an eigenvalue −1 + cW (0) ≈ −1 − 0.177c. This eigen- value will be negative for all c. When c = 5 the most excitable mode has eigenvalue −1 + 5W (kmax) ≈ 0.1, and when c = 25 the most excitable mode has eigenvalue −1 + 25W (kmax) = 4.5. Even though these eigenval- ues are greater than 0 we see decay both when c = 5 and when c = 25, and we do not see the effect of the small initial perturbation away from the constant value Y = 0.5 in time t = 0.5. Here the damping effect is greater than the exponential in (12) initially, although eventually the exponential causes the pattern to appear 10 is indeed the case (Figure 5a). When a small amount of i.i.d. Gaussian noise, σ = 0.5, was added on each iteration in (19), however, a spatial pattern is evident (Figure 5b). More noise, σ = 1.0, makes the pattern more apparent (Figure 5c). Thus, again we verify that noise can both reveal weak, otherwise initially (at short time intervals) unobservable, spatial patterns, and also sustain them at observable levels across time. The dominant wave num- ber (in the sense defined earlier) of the spatial pattern does not depend on the standard deviation of the noise, σ. It will depend, however, on the standard deviation of the smoothing kernel, η, as predicted by the maximum R(k) in Figure 1. We explore this relation in of (1 − cW (k)) the next section. 3. Coupling and noise smoothing As discussed earlier, noise smoothing can be expected to affect the spatial pattern created by the Mexican Hat coupling. When the maximum λk > 0, as here, (12) ap- plies. The interaction of c and η in (12) is still controlled by the ratio of σ2 k/λk, so we expect the effects of the noise smoothing to be consistent with those for the case where all λk < 0. These effects are displayed in Figure 6 for c = 22.5, σ = 1.0 and values of η = 0.15, 0.5, 1.3. Figure 5c shows the results of a simulation with these same parameter values but no noise smoothing. When η = 0.15 < 0.28, as pre- dicted by (15), the variance should be increased slightly but the dominant mode is not affected, and this is appar- ent in Figure 6a. When η = 1.3 the variance is much re- duced but the dominant mode is shifted toward 0, reduc- ing by one the number of cycles induced by the Mexican Hat coupling alone (Figure 6c). When η = 0.5 > 0.28, however, the dominant mode is unaffected and the vari- ance also reduced, creating a more apparent spatial pat- tern, the 'best' in our collection of solutions (Figure 6b). The interaction between these various factors in their ef- fects on the spatial pattern induced by the Mexican Hat coupling thus creates optimal combinations of c, η for the emergence of the spatial pattern. This phenomenon of an optimal pair (c, η) is remi- niscent of stochastic resonance (or stochastic facilitation [25]), in which tuning of the noise strength and threshold yields optimum performance. Of course, the greater c is, the stronger the pattern, so this analogy only applies for situations where damping is sufficient that the pattern is only revealed and sustained by optimally smoothed noise. Smoothing that is too broad imposes a lower frequency on the array, and interferes somewhat with the pattern created by the Mexican Hat operator, as seen in the case where c = 22.5, η = 1.3. FIG. 3. Spatial patterns in the case λk < 0 over interval t = 0 to t = 25 for i.i.d. noise (a) and for smoothed noise with η = 0.5 (b). Here ∆t = 0.0025, c = 4.5, σ = 1.0. Top row: amplitude Y (t, x) ; second row Y (t, x) at t = 25; third row: F ; bottom row: FFT amplitude. Here the expected number of cycles in L/2π is again 8, as described in the text. (not shown). When c = 75, however, the most excitable mode has eigenvalue −1 + 75W (kmax) ≈ 15.5. Moreover, again, arg max{W (2πk/L : k ∈ Z≥0} ≈ (L/2π) × 2 ≈ 8, and so (with no noise) the most excitable mode should have period 8. By time t = 0.5 we begin to see the expo- nential growth and the pattern of 8 spatial cycles for the Yj(t) amplitude, FFT amplitude, and F , produced by the Mexican Hat operator. See the top graph of Figure 4c. 2. I.I.D. noise Figure 5 displays the effects of adding i.i.d. Gaussian noise to the neural field equation with the same param- eters for w(x) as in Figure 4 except that in Figure 5 c = 22.5 so λk > 0 for some k. We would expect, based on Figure 4b,where c = 25, that little or no indication of spatial pattern would be apparent when σ = 0, and that η=0.5&i.i.d.&FFT&Amplitude&FFT&Amplitude&(a) i.i.d. (b) η =0.5 11 FIG. 4. Damping of spatial patterns for which cW (kmax) > 1. Top row: amplitude Y (t, x) ; middle row: F ; bottom row: FFT amplitude. As c (indicated at top of figure) increases the spatial pattern becomes more apparent. Here σ = 0 and thus noise smoothing is irrelevant. Here kmax = 2.0263 and the expected number of cycles in L/2π is 8, as described in the text. C. Spatial smoothing of noise without coupling As described earlier, there are spatial modes of the smoothed noise itself. For small η these are spread over a range of values of k, whereas for large η they are con- centrated near k = 0, as shown in Figure 1, although the dominant mode is always 0. Because the dominant mode of 0 creates no spatial pattern, however, we might expect that when there are non-zero spatial modes of significant amplitude, the noise alone, in the absence of Mexican Hat coupling, might induce a spatial pattern. Figure 7 displays solutions to the stochastic neural field equation (2) with c = 0 and Gaussian-smoothed noise. That is, there is no coupling via the Mexican Hat opera- tor. In these cases, however, η has been varied, from i.i.d. spatio-temporal noise to η = 0.67 and to η = 1.3. We observe that smoothing the noise itself creates a spatial pattern with FFT amplitude depending on η, effecting what we could term a 'coupling through partially shared noise.' The FFT amplitude decays exponentially toward the higher frequencies, as expected. With η = 2.4 the FFT amplitude is concentrated close to spatial frequency 0 (not shown). Indeed, when the Gaussian smoother has significant weighting over the entire ring, η > 2, there is only one large bump in the typical simulated path, and the FFT amplitude is concentrated at a spatial frequency of one cycle over L (not shown). Finally, the spatial pat- tern created by the smoothing kernel can be expected to interact with that created by the Mexican Hat operator to create a sustained and powerful standing wave when the noise smoothing is optimal, as in Figure 6b when η = 0.5 or to overwhelm the Mexican Hat pattern when noise smoothing is too great, as in Figure 6c when η = 1.3 . Pattern produced by Gaussian noise smoothing is, however, significantly different from pattern produced by Mexican Hat coupling. Gaussian noise smoothing with c = 0 tends to produce stochastic paths that resem- ble irregular bumps over space, and these bumps move around in the spatial array as the field evolves, making the evolving neural field resemble a chimera of emerging and fading pattern [26]. The Mexican Hat coupling with or without smoothing tends to result in stochastic paths that resemble stripes rather than bumps (Figures 5, 6), although the stripes do move around somewhat as the field evolves, especially when the noise is not optimally smoothed. VI. DISCUSSION We have illustrated, in the context of a "standard" [7] stochastic neural field equation, (2), that a Mexican Hat convolution kernel produces spatial patterns that (even- tually) can be revealed and sustained by noise, even when all eigenvalues are negative, whereas without noise the damping tends to dominate the pattern. Moreover, over long time intervals, Gaussian-smoothed noise alone also c=5$c=25$c=75$FFT$Amplitude$FFT$Amplitude$FFT$Amplitude$(a) (c) (b) 12 FIG. 5. Added noise reveals and sustains the spatial pattern. Rows are the same as in Figure 4, noise level σ is indicated at the top of each column. Here c = 22.5, with no noise smoothing. produces a spatial pattern, and the two sources of pattern interact. It has been known for some time that a Mexican Hat convolution kernel produces spatial patterns, similar to Turing patterns, in a variety of contexts (e.g., [22, 24]). It was known previously that noise can reveal and sus- tain such patterns which are otherwise damped, much as quasicycles are revealed and sustained by noise in tem- poral stochastic processes. We studied major features of this process, exploring the dependence of the pattern on the strength of the Mexican Hat coupling and the width of the noise smoothing. First, we found a parametric measure of the interaction between noise sharing/smoothing and Mexican Hat cou- pling, based on the Fourier expansion of the neural field equation. We then demonstrated, for the case where all eigenvalues of the operator are negative, a close corre- spndence between the predictions of the continuous the- ory and an Euler-Maruyama discretization of the the- ory. Next we explored, for the case where at least one eigenvalue of the operator is positive, the relationship be- tween the coupling strength, c and the width of the noise smoothing, η. We found a family of 'best combinations' of parameters controlling coupling and noise smoothing that produced the strongest patterns. A particular find- ing is that Gaussian noise smoothing can itself, without coupling, produce spatial patterns in the context of neu- ral field equations, and likely in other contexts as well. A. Development of spatial pattern in time We explored solutions to (2) and (19) over both short, t = 0.5, and longer, t = 25, time intervals. The re- sults over the longer time intervals demonstrate that even when all eigenvalues are negative the Mexican Hat coupling can produce spatial patterns in the presence of added noise, and these patterns can be quite clear when noise is shared/smoothed over a local neighbourhood. The brain, however, does not remain in a single state for such long time intervals. A more likely scenario for application of our results to real brains is that the brain's state changes over shorter time intervals, typically every few hundred milliseconds or faster. Thus, our results for the shorter time intervals are probably most relevant. Importantly, it would be inefficient for the brain to have evolved a system of local coupling, the Mexican Hat cou- pling, for which all eigenvalues are negative, so that very long time intervals are required for the coupling to cre- σ=0$σ=1.0$σ=0.5$FFT$Amplitude$FFT$Amplitude$FFT$Amplitude$(a) (c) (b) 13 FIG. 6. Smoothed noise reveals and sustains the spatial pattern as well as making it more regular. Rows are the same as in Figure 4 except for addition of row 2, the amplitude after the 10,000th iteration. Smoothing kernel standard deviation, η, is indicated at the top of each column. Here the standard deviation of the Gaussian noise is σ = 1.0, and the coupling strength is c = 22.5. The two somewhat contrasting effects of the smoothed noise interact with the spatial pattern created by the Mexican Hat operator. ate spatial patterns. To illustrate this, compare Figure 2E with Figure 6 middle column. The FFT amplitudes of the spatial patterns are roughly equivalent in the two cases. The time required to reach this state, however, is t = 25 for Figure 2E, where c = 4.5 and λk < 0 for all k, whereas the time required to reach the state in Figure 6 is t = 0.5, where c = 22.5 and λk > 0 for at least one k. Thus, even though we don't know in terms of scaling how this space-time model might relate to real systems, the scenario where at least one λk > 0 is the most likely to be present. Combined with optimal noise sharing, e.g., with η = 0.5 in our simulations and in the example just described, this scenario would implicate a functional role for the local Mexican Hat coupling, such as enhancing edges of neural representations of visual stimuli. produce spatial patterns has, we believe, been unappre- ciated until now. In fact, the Fourier transform of the process (2) predicts its spatial modes, including the case in which, because c = 0, the Mexican Hat kernel has no effect. In this case, where the smoothing kernel acts alone, a greater spread of noise smoothing leads to a nar- rower range of k with significant power. Smoothing that is significant over the entire lattice leads to a range of k that is close to 0. The pattern that results from the in- teraction of the Mexican Hat coupling and the coupling by smoothing of noise is a combination of their respec- tive spatial modes, as reflected in Figure 1. Which modes dominate in a particular implementation depends on the weighting of the respective operators and their extents with respect to the size of the system. B. Spatially-smoothed, noise-induced patterns C. Relation to other work The fact that spatially-smoothed noise, i.e. noise that has non-zero correlation length in space, can by itself There is an extensive literature on stochastic neural field equations. We have chosen a representative, as- η=0.15'η=1.3'η=0.5'FFT'Amplitude'FFT'Amplitude'FFT'Amplitude'(a) (c) (b) 14 FIG. 7. Applying the Gaussian noise-smoothing kernel across neighbors creates a spatial pattern. Rows are the same as in Figure 4. Here c = 0 (so no coupling via Mexican Hat operator), σ = 0.5, and standard deviations of the Gaussian smoothing kernel, η, are indicated at top of figure and i.i.d. indicates no noise smoothing. sorted, sample from this literature and point out similar- ities and differences to the present work, and directions for future studies. Meyer, Ladenbauer and Obermayer [2] produced a grid array of spiking neurons with a Mexican Hat coupling structure and measured the covariance pattern of the resulting spike counts while varying the parameters of the Mexican Hat coupling. They found an oscillating pattern of correlation decay around particular fixed neu- rons, produced by Mexican Hat coupling, with wider in- hibitory than excitatory coupling. This pattern did not appear when an inverse Mexican Hat coupling, in which the excitatory coupling extended further than did the inhibitory coupling, was imposed on the grid. The sys- tem was driven by external i.i.d. synaptic noise to each neuron, so that the output spike pattern correlation was necessarily produced by the coupling. The approach of the study [2] was complimentary, or dual, to the approach and objective of the present paper. In [17, 27] there is a cubic reaction term that suc- ceeds in keeping the process stochastically bounded. This is different from the thresholded identity function used in the present work. In [27] the coupling operator is (1 + (∂2/∂x2))2, which has an effect similar to a Mexican Hat, whereas in [17] an effectively Mexican Hat opera- tor, written differently, is used. Both in [27] and in [17] the noise is either uncorrelated spatially or, the other ex- treme, 'global fluctuations,' in which the same noise is added to all components of the neural field at each time point. This is in contrast to locally spatially smoothed noise, which we studied here. In these papers the an- alytic method of center manifold theory together with adiabatic elimination is used to obtain solutions to the neural field equation. The paper [24] studied a model that creates a mov- ing front between states 0 and 1 using a cubic reaction term as in [27]. At the same time a Mexican Hat kernel together with a diffusion term creates a Turing pattern. Homogeneoous solutions coexist with spatially periodic states. There is no stochastic term, however, and the effects of noise in this model are unknown. In [8], Gaussian white noise, as in [28] together with o + ∇2)2 and, again, a spatial coupling of the form (K 2 cubic reaction term, in a Stratonovich SDE, create pat- terns in R2. These patterns resemble various highly reg- ular patterns of vegetation that occur on slopes in semi- deserts around the world. In Touboul's paper [9] space-dependent delays are in- troduced. Again the noise is not smoothed across space. A relevant result is convergence-in-law of network equa- tions. These are in continuous time and discrete in num- bers of neurons and of populations, both of which in- η=0.001&η=0.1&η=0.2&!!!!!!i.i.d.!!!!!!!η=1.3!!!!!!!η=0.67!FFT!Amplitude!FFT!Amplitude!FFT!Amplitude!(a) i.i.d. (c) (b) crease to infinity, the 'neural-field limit.' The limit is a particular McKean-Vlasov equation, a stochastic neural mean field equation with delays. These other works suggest various additional ways to pursue the questions studied here. For example, one could insert a cubic reaction term in (2) instead of us- ing the function S to maintain stochastic boundedness of the process. Additional analytic results might well be obtained using the center manifold theory for the case k = 0 as in [17, 27]. Finally, the present paper can also be seen in the context of the broader field of pattern formation aris- ing from stochastic differential equations. A fairly re- cent text summarizing many problems and results in this field is that of Blomker [29]. In that work methods are described for the approximation of stochastic par- tial differential equations near a change of stability using amplitude equations. Blomker focuses on rigorous er- ror estimates for such approximations with the aim of enabling their use in physics and applied mathematics. This text provides many useful clues about how to ex- tend the present work, including a detailed description of approximative center manifold theory. 15 Appendix A Proposition: Consider the processes Bk = {Bk(t) : t ≥ 0} defined in (10) for non-negative integers k. (i) The processes B0, B1, B2, . . . are independent. (ii) B0 is scalar Brownian motion with variance LR(0). (iii) For k ≥ 1, Bk is 2-dimensional Brownian motion with variance (L/2)R(2πk/L). Proof: The family {Bk(t) : k ≥ 0, t ≥ 0} is a complex valued mean 0 Gaussian process, so it enough to calculate covariances. Each process Bk is a complex-valued Gaus- sian process with independent increments in time, so it is enough to calculate covariances of the complex random variables {Bk(1) : k ≥ 0}. Write Bk(1) = B1 k. Since r(x) = r(−x) we have k +iB2 (cos 2πkx/L)r(x) dx = R(2πk/L), (sin 2πkx/L)r(x) dx = 0. 0 We calculate some covariances for k, (cid:96) ≥ 0. First the real parts: (cid:90) L (cid:90) L 0 kB1 (cid:96) = E E(cid:2)B1 (cid:3) (cid:34)(cid:32)(cid:90) L (cid:90) L (cid:90) L (cid:90) L = 0 0 0 = cos 0 = R(2π(cid:96)/L) = R(2π(cid:96)/L) kB2 (cid:96) = E E(cid:2)B2 (cid:3) (cid:34)(cid:32)(cid:90) L (cid:90) L (cid:90) L (cid:90) L = 0 0 0 = sin 0 = R(2π(cid:96)/L) = R(2π(cid:96)/L) Similarly for the imaginary parts: (cid:33)(cid:32)(cid:90) L (cos 2πkx/L)G(1, x) dx (cid:33)(cid:35) (cos 2π(cid:96)y/L)G(1, y) dy 0 (cos 2πkx/L)(cos 2π(cid:96)y/L)r(x − y) dxdy 2π(cid:96)(y − x) L − sin 2π(cid:96)x L sin 2π(cid:96)(y − x) L cos 2πkx L 0 2π(cid:96)x (cid:32)(cid:90) L (cid:0) cos (cid:90) L  L if k = (cid:96) = 0 L/2 if (cid:96) = k else. L 0 0 (cos 2πkx/L)(cos 2π(cid:96)x/L) dx (cid:33)(cid:32)(cid:90) L (cid:33)(cid:35) (sin 2π(cid:96)y/L)G(1, y) dy 0 2π(cid:96)(y − x) L cos + cos 2π(cid:96)x L sin 2π(cid:96)(y − x) L (sin 2πkx/L)(sin 2π(cid:96)y/L)r(x − y) dxdy (sin 2πkx/L)G(1, x) dx 2πkx L 2π(cid:96)x (cid:0) sin (cid:32)(cid:90) L (cid:90) L (cid:26) L/2 if (cid:96) = k (cid:54)= 0 L 0 0 0 else. (sin 2πkx/L)(sin 2π(cid:96)x/L) dx (cid:33) (cid:1)r(x − y) dy dx (cid:33) (cid:1)r(x − y) dy dx Finally for the 'mixed' terms: kB2 (cid:96) = E (cid:3) E(cid:2)B1 (cid:34)(cid:32)(cid:90) L (cid:90) L (cid:90) L (cid:90) L = 0 0 = cos 0 0 (cid:32)(cid:90) L (cid:90) L 0 L 2πkx (cid:0) sin = R(2π(cid:96)/L) = 0. 0 (cos 2πkx/L)(sin 2π(cid:96)x/L) dx (cid:33)(cid:32)(cid:90) L (cid:33)(cid:35) (sin 2π(cid:96)y/L)G(1, y) dy 0 (cos 2πkx/L)G(1, x) dx (cos 2πkx/L)(sin 2π(cid:96)y/L)r(x − y) dxdy 2π(cid:96)x L cos 2π(cid:96)(y − x) L + cos 2π(cid:96)x L sin 2π(cid:96)(y − x) L 16 (cid:33) (cid:1)r(x − y) dy dx Since orthogonality implies independence for Gaussian random variables, the results follow directly. COMPETING INTERESTS AUTHOR'S CONTRIBUTIONS All authors contributed to the conceptualization and writing of the paper. The numerical simulations were accomplished by LMW. ACKNOWLEDGEMENTS The authors declare that they have no competing in- Lawrence M. Ward was supported by Discovery Grant terests. A9958 from NSERC of Canada. [1] B. Doiron, A. Litwin-Kumar, R. Rosenbaum, and K. Josic, Nature Neuroscience G.K. Ocker, 19, 383 (2016). [2] R. Meyer, J. Ladenbauer, and K. Obermeyer, Frontiers in Computational Neuroscience 11, 34 (2017). [3] A. Hutt, M. Bestehorn, and T. Wenneker, Network: Computation in Neural Systems 14, 351 (2003). [4] T. Butler and N. Goldenfeld, Physical Review E 80, 030902R (2009). [14] M. Ferrante and M. Sanz-Sol´e, ESAIM Probability and Statistics 10, 380 (2006). [15] M. Sanz-Sol´e and M. Sarr`a, in Seminar on Stochastic Analysis, Random Fields and Applications, III (Ascona, 1999), Progress in Probability, Vol. 52 (Birkhauser, Basel, 2002) pp. 259 -- 268. [16] C.W. Gardiner, Handbook of stochastic methods for physics, chemistry and the natural sciences 2nd ed (Springer, N.Y., 1990). [5] T. Butler and N. Goldenfeld, Physical Review E 84, [17] A. Hutt, A. Longtin, and L. Schimansky-Geier, Physica 011112 (2011). D 237, 755 (2008). [6] A. McKane, T. Biancalani, and T. Rogers, Bulletin of [18] P. Boxler, Probability Theory and Related Fields 83, Mathematical Biology 76, 895 (2014). 509-545 (1989). [7] O. Faugeras and J. Inglis, Journal of Mathematical Biol- ogy 71, 259 (2015). [8] S. Scarsaglio, F. Laio, P. D'Odorico, and L. Ridolfi, Mathematical Biosciences 229, 174 (2011). [9] J. Touboul, The Annals of Applied Probability 24, 1298 (2014). [19] M.S. Nixon and A.S. Aguado,Feature Extraction and Im- age Processing for Computer Vision: Third Edition (El- sevier, 2008). [20] D. Marr, Vision (The MIT Press, 1982). [21] P. C. Bressloff, Journal of Physics A : Mathematical and Theoretical 45, 033001 (2012). [10] S. Song, P.J.Sjostrom, M. Reigl, S. Nelson, and [22] J. D. Murray, Mathematical Biology (Springer, Berlin, D. Chklovskii, PLoS Biology 3(3), e68 (2005). 1989). [11] B. Haider and D. McCormick, Neuron 62, 171 (2009). [12] A. A. Faisal, L. P. J. Selen, and D. Wolpert, Nature Reviews Neuroscience 9, 292 (2008). [23] P. E. Kloeden and E. Platen, Numerical Solution of Stochastic Differential Equations (Springer-Verlag, Berlin, 1992). [13] R. Dalang and N. Frangos, The Annals of Probability 26, [24] J. Siebert and E. Scholl, Europhysics Letters 109, 40014 187 (1998). (2015). [25] M. D. McDonnell and L. M. Ward, Nature Reviews Neu- Berlin, 1986) pp. 265 -- 439. roscience 12, 415 (2011). [26] Y. Kuramoto and D. Battogtokh, Nonlinear Phenomena in Complex Systems 5, 380 (2002). [27] A. Hutt, A. Longtin, and L. Schimansky-Geier, Physical Review Letters 98, 230601 (2007). [28] J. Walsh, in ´Ecole d'´et´e de probabilit´es de Saint-Flour, XIV -- 1984, Lecture Notes in Mathematics. (Springer, [29] D. Blomker, Amplitude Equations for Stochastic Par- tial Differential Equations (World Scientific, Singapore, 2007) . 17
1709.03000
3
1709
2018-10-30T15:34:12
Network constraints on learnability of probabilistic motor sequences
[ "q-bio.NC" ]
Human learners are adept at grasping the complex relationships underlying incoming sequential input. In the present work, we formalize complex relationships as graph structures derived from temporal associations in motor sequences. Next, we explore the extent to which learners are sensitive to key variations in the topological properties inherent to those graph structures. Participants performed a probabilistic motor sequence task in which the order of button presses was determined by the traversal of graphs with modular, lattice-like, or random organization. Graph nodes each represented a unique button press and edges represented a transition between button presses. Results indicate that learning, indexed here by participants' response times, was strongly mediated by the graph's meso-scale organization, with modular graphs being associated with shorter response times than random and lattice graphs. Moreover, variations in a node's number of connections (degree) and a node's role in mediating long-distance communication (betweenness centrality) impacted graph learning, even after accounting for level of practice on that node. These results demonstrate that the graph architecture underlying temporal sequences of stimuli fundamentally constrains learning, and moreover that tools from network science provide a valuable framework for assessing how learners encode complex, temporally structured information.
q-bio.NC
q-bio
Network constraints on learnability of probabilistic motor sequences Ari E. Kahn1,2,3, Elisabeth A. Karuza4, Jean M. Vettel3,5,2, and Danielle S. Bassett2,6,7,8* 1Department of Neuroscience, University of Pennsylvania, Philadelphia, PA 19104 USA 2Department of Bioengineering, University of Pennsylvania, Philadelphia, PA 19104 USA 3Human Research and Engineering Directorate, U.S. Army Research Laboratory, Aberdeen, MD 21001 USA 4Department of Psychology, University of Pennsylvania, Philadelphia, PA 19104 USA 5Department of Psychological and Brain Sciences, University of California, Santa Barbara, CA 93106 USA 6Department of Electrical & Systems Engineering, University of Pennsylvania, Philadelphia, PA 19104 USA 7Department of Neurology, Perelman School of Medicine, University of Pennsylvania, Philadelphia, PA 19104 USA 8Department of Physics & Astronomy, College of Arts and Sciences, University of Pennsylvania, Philadelphia, PA 19104 USA *To whom correspondence should be addressed: [email protected]. 8 1 0 2 t c O 0 3 ] . C N o i b - q [ 3 v 0 0 0 3 0 . 9 0 7 1 : v i X r a Keywords: graph learning, statistical learning, motor sequence learning, graph theory, probabilistic sequences Abstract Human learners are adept at grasping the complex relationships underlying incoming sequential input1. In the present work, we formalize complex relationships as graph structures2 derived from temporal associations3,4 in motor sequences. Next, we explore the extent to which learners are sensitive to key variations in the topological properties5 inherent to those graph structures. Participants performed a probabilistic motor sequence task in which the order of button presses was determined by the traversal of graphs with modular, lattice-like, or random organization. Graph nodes each represented a unique button press and edges represented a transition between button presses. Results indicate that learning, indexed here by participants' response times, was strongly mediated by the graph's meso-scale organization, with modular graphs being associated with shorter response times than random and lattice graphs. Moreover, variations in a node's number of connections (degree) and a node's role in mediating long-distance communication (betweenness centrality) impacted graph learning, even after accounting for level of practice on that node. These results demonstrate that the graph architecture underlying temporal sequences of stimuli fundamentally constrains learning, and moreover that tools from network science provide a valuable framework for assessing how learners encode complex, temporally structured information. Main Our ability to interact with our environment necessitates that we parse complex stimuli into smaller units, such as words and phrases in language input, or events in streams of visual stimuli. This essential process relies at least in part on the statistical regularities present around us, and often operates automatically and without any explicit, verbalizable knowledge of underlying rules1. Statistical regularities can be inferred from various sources of information, including but not limited to the temporal order in which stimuli are experienced. As early as infancy, humans reliably detect the probabilities with which one stimulus transitions to another (transition probabilities, such as one syllable following another in spoken language) and the frequencies with which stimuli temporally co-occur (co-occurrence frequencies)6. Similar forms of pattern sensitivity have been observed beyond the language domain, including motor learning7,8 and visual event segmentation9,10. Ongoing research examines which types of statistics induce learning -- including statistical associations between movements11 and between non-linguistic sounds12. Moreover, evidence suggests that, depending on context, learners can extract both adjacent 2 and non-adjacent dependencies between stimuli13,14. Taken together, these studies suggest that second- or third-order statistical relationships may be encoded implicitly, and furthermore, that higher-level organizational principles themselves might be implicitly learned. Indeed, recent work has shown that temporal ordering of visual stimuli can convey the organizational principle of modularity3. This observation opens up the possibility of studying whether certain organizational principles are more or less facilitative of learning, and whether information embedded in certain organizational structures might be easier to learn than information embedded in others. An ideally suited language in which to define such higher-order principles is network science2, an emerging interdisciplinary field that addresses the architecture, dynamics, and design of complex systems composed of many connected parts5. The set of parts (network nodes) and connections (network edges) are often parsimoniously encoded in a mathematical object called a graph15. In the context of learning, we can construct a graph that encodes the pattern of relationships between objects, movements, or sounds. Prior theoretical work16 has addressed the relationship between graph structure and artificial grammars (such as in Cleeremans & McClelland 11), and we build on this work by empirically addressing the impact of graph-based properties. Recently Karuza et al. 4 capitalized on this approach to define a graph from which the temporal ordering of visual stimuli was drawn. Learners exhibited a robust on-line measure of learned graph structure: a surprisal effect defined as an increase in reaction time when transitioning between modules. Importantly, this surprisal effect was dependent on the type of traversal through the graph, and was more strongly pronounced when traversals through the graph provided redundancy in local information. While the manner in which a graph is traversed can influence learning, the nature of the graph itself may serve as an even more fundamental constraint on the potential for humans to learn organizational principles of information. Many real-world systems including language17, conceptual knowledge18,19, social groups20, and societies21 display non-trivial higher-order structure such as clustering or hub-and-spoke architecture that is relevant for how humans can optimally communicate, interact, and ultimately survive in their environment. Moreover, our knowledge about these systems unfolds and grows over time as we experience new parts (nodes) and their relations (edges). Understanding the impact of such higher-order structure on learning could help to explain why knowledge of some (natural or man-made) systems may be more easily acquired than others, and why individuals differ in their capacity to learn them. It may also shed light on the question of how humans can generalize from local statistical information to develop representations of broad-scale organizational patterns22. Here, we examined whether the higher-order regularities of three graph structures influenced implicit learning of statistical relationships among temporally ordered stimuli. Specifically, we trained subjects on a self-paced Serial Reaction Time (SRT) 3/29 task, where each trial was drawn from a traversal through a graph. Each node represented a stimulus, and each edge represented a possible transition between two nodes. Based on prior work demonstrating learners' sensitivity to higher-order statistics in SRT-like tasks11, we hypothesized that learners would display sensitivity to graph structure as evidenced by a surprisal effect. Next, we systematically varied graph structure to examine the impact of graph topology on the acquisition of complex, multi-element motor sequences. We hypothesized that learners would display increasingly rapid execution of button presses when presented with modular graph structures in comparison to either random or lattice graphs. The predicted preference for modular graphs is based on evidence across disparate fields of scientific inquiry. Modular topologies are more frequently observed in real-world systems than either random or lattice-like topologies23. Moreover, the clustering and hierarchy of modular graphs in natural systems can emerge in response to constraints on network wiring costs24, and similar constraints on complexity may impact learnability itself. Indeed, we predicted that learning mechanisms should be tuned to statistical features of natural stimuli25. Finally, we hypothesized that learning performance would vary over different regions of the graph based on both local and global properties. Local properties are those inherent in a single node and its immediate neighbors, and global properties encompass organizational features of the entire graph such as clustering or repeated structure. Together these properties can be used to assess learners' sensitivity to variations in the graph across topological scales. Collectively, the results we report below suggest that learners extract and exploit the graph topology defining temporal sequences of stimuli, and that topological features impact speed of acquisition. Results Setup: We recruited 381 unique participants: 109 participants for a first experiment, 59 participants for a second experiment, and 223 participants for a third experiment. Subjects performed a self-paced SRT motor response task using a keyboard. Stimuli were five grey squares in a horizontal row; to indicate a target key or pair of keys that the subject was meant to press, the corresponding square(s) would be outlined in red (Fig. 1A). The squares corresponded spatially with keys 'Space', 'H', 'J', 'K' and 'L', such that the left square represented 'Space' and the right square represented 'L' (Fig. 1B). These keys were chosen so that the subject's hand could ergonomically rest over all five keys at once. Sequence Generation: The order in which stimuli were presented to the subject was prescribed by either a random or a Hamiltonian walk on a graph of N = 15 nodes connected by E = 30 edges (Supplementary Table 1). Random Walk: For any two nodes that were not connected by an edge, the transition probability was equal to zero. For any two nodes that were connected by an edge, the transition probability was equal to 1 divided by the number of edges emanating from the pre-transition node. Hamiltonian Walk: A walk composed of 4/29 Figure 1. Experimental Setup. (A) An example of the first few steps of a graph traversal defined by a walk on the graph. Top: Each node is uniquely associated with a key combination, and the sequence of key combinations is determined by a walk on the graph. Bottom: A series of trials are presented to the participant. The red squares indicate which keys to press on that trial. Colored arrows illustrate the edge from the graph at top being traversed. However, the participant only is shown the five squares. (B) The mapping between fingers and keys, and average reaction times for each key press. Top: A schematic of the mapping between visual stimuli (squares) and response effectors (fingers). Bottom: The average reaction time (RT) for each key or pair of keys across all data. The diagonal elements of the matrix represent trials in which a single key was pressed, and the off-diagonal elements of the matrix represent trials in which a pair of keys was pressed. (C) The three graph structures that we examine in this study. From left to right, we show a modular graph, a lattice graph, and a random graph with N = 15 nodes connected by E = 30 edges. a series of cycles, each of which visited every node on the graph exactly once. Each cycle was randomly generated starting from a node adjacent from the endpoint of the previous cycle. Graphs: We compared learning rates across 3 different graph topologies, each consisting of 15 nodes and 30 edges: a modular graph, a lattice graph, and a random graph (Fig.1C). Briefly, 5/29 HJKLSpaceAC721859107410739647658971043109480494611818471036809LKHJSpaceFinger 1RT (ms)HKJLSpaceB1200950700TrialNode3143512345. . .. . .51423 the modular graph consists of three clusters of five interconnected nodes. The lattice can be thought of as a grid, wrapping around at its boundary. The random graphs differed between subjects and had no consistent organizational principles besides constituting a single connected component and maintaining the same number of nodes and edges as the other two graphs. (See Methods for formal definitions.) For both the modular and lattice graphs, the equal degree distribution coupled with a random walk leads to uniform pairwise probabilities for all possible transitions from a given node in the graph. Experiment: We ran three experiments, each consisting of two back-to-back stages that differed in which graph and walk type was used to generate the stimulus sequence. The first experiment considered learning on one of three distinct topologies (modular, lattice, or random) instantiated as sparse graphs containing only a minority of possible edges between nodes; learning from one of the structured topologies was followed by learning from a fully connected graph structure, in an effort to identify any changes in learning driven by the addition of novel edges. Pragmatically, the experiment was operationalized with a within-subjects design. Specifically, the first stage of the experiment used either a modular, lattice, or random graph to generate a sequence of 1500 stimuli via a random walk on the graph. The second stage of the experiment used a fully connected graph to generate a sequence of 500 stimuli via a random walk on the graph (which amounted to a random stimulus order). With every node connected to every other node (eliminating any informative structure of the input), the fully connected graph allowed a comparison between previously trained edges and novel edges. The second experiment consisted of 1500 trials of a random walk on a modular graph, followed by 500 trials of a Hamiltonian walk on the same graph, allowing us to confirm that learned differences transferred to a different walk type on the same graph structure. The third experiment directly compared learning effects between graph types, accounting for individual variability in baseline reaction times and learning rates using a within-subjects design. Similar to the first experiment, the first stage of the experiment consisted of a sequence of 1500 stimuli via a random walk on either a modular graph, a lattice graph, or a random graph. However, unlike the first experiment, the second stage was another sequence of 1500 stimuli via a random walk on one of the remaining two sparse structured graph types. Analysis: We verified the surprisal effect using all modular graph traversals from stage 1 (Experiment 1). The effect of modifying the graph topology through the addition of novel edges was assessed using a sparse graph followed by a full graph in a between-subjects design (Experiment 1). Our assessment of the sensitivity of the surprisal effect to a switch from a random walk to a Hamiltonian walk was based on Experiment 2. Our comparison of behavioral performance between the sparse graph types (modular, lattice, and random) was based on a within-subjects design (Experiment 3). Finally, our comparison of local and global graph properties was based on data acquired during the random walk traversal of a sparse random graph (stage 1 of Experiments 1 and 3). After determining differences in reaction time by key or key combination (Fig. 1B) (which was then added as a regressor in 6/29 all subsequent models; see Methods), we asked whether participants displayed evidence of learning the probabilistic motor sequence. A commonly studied marker of motor skill learning is an exponential drop-off in movement time with trials practiced26. We observed this drop-off, despite our task being probabilistic rather than deterministic. Learners exhibited large decreases in reaction time on average over the course of the first experiment. Across the initial structured graph stage, we observed mean reaction time decreasing by nearly 500 ms (Fig. 2A, black line. See Supplementary Figures 1 & 2 for additional information). The observed overall decrease in reaction time with training suggested general improvement in task performance. To explicitly evaluate learning based on graph structure, we tested for the cross-cluster surprisal effect, a previously reported measure of graph learning on modular graphs4. Briefly, the cross-cluster surprisal effect is a significant increase in reaction time when a participant traverses an edge located between clusters, in comparison to when a participant traverses edges within clusters. Using a mixed effects model (see Methods), we observed a statistically significant increase in reaction time across edges that connected two clusters (Fig. 2A,B; linear mixed effects model; t(29) = 3.61, p < 0.002, expected increase of 63.6 ms; 95% confidence interval: 29.07 to 98.10, see Supplementary Table 2). To provide an intuitive visualization of this finding, we first note that the graph was symmetric across the three clusters; because the starting position and traversal direction within the graph varied for each subject, the distinction between the three clusters was arbitrary when comparing across subjects. We therefore remapped each edge to the equivalent edge within a single canonical cluster, which visually highlights the clear difference in reaction times for between- versus within-cluster edges (Fig. 2C). The observed surprisal effect suggests that participants are sensitive to graph structure. However, an alternative explanation is that the surprisal effect reflects a difference in processing cost inherent to local repetitions associated with a random walk on the modular graph. To either support or dismiss this alternative explanation, we asked whether the surprisal effect would persist when we modified the walk to sample sparsely from each module in time, eliminating within-module repetitions of button presses. Prior work4 suggests that learners are unlikely to show a surprisal effect when exposed to a Hamiltonian walk, where each node is only visited once per cycle. However, if clusters correspond to a learned feature of the graph, then we expect that learners first trained on a random walk (where we expect a surprisal effect) on a graph, and then exposed to a Hamiltonian walk (where we do not expect a surprisal effect) on the same graph, will continue to show the surprisal effect. In this experiment, subjects were first exposed to 1500 trials of a random walk on the modular graph, followed by 500 trials of a Hamiltonian walk on the same graph. We were interested to determine whether a subject's surprisal effect in the first stage was correlated with their surprisal effect in the second stage. Using a mixed effects model, we found that our estimate for the 7/29 Figure 2. Modular Graph Learning Effects. (A) Mean reaction times (RTs) as a function of trial number for stage 1 of Experiment 1, among participants exposed to the modular graph. The red line indicates the mean for cross-cluster trials, and the black line indicates the mean for all other trials, each binned in sets of 30 trials (n=30 subjects). (B) Mean reaction times on correct trials for the modular graph. An increase in reaction time across cluster boundaries can be seen, here visualized by yellower colors in the matrix elements that sit between the larger blocks. (C) Mean reaction times collapsed across the symmetric structure of the modular graph. All three clusters were structurally identical and starting position was randomized between subjects, so we combine reaction times across the three clusters into one 'canonical' cluster for visualization purposes only. The mean increase in reaction time between clusters is more apparent, here visualized by yellower colors on the edges that connect the top cluster with the two bottom clusters. (D) Relationship between surprisal effect on stage 1 (random walk) and surprisal effect on stage 2 (Hamiltonian walk) for each subject of Experiment 2. Subjects that displayed a strong surprisal effect in stage 1 likewise do so when the walk structure is changed (n=59). surprisal effect was significantly reduced and not significant in the Hamiltonian walk, with an estimated increase in RT of 7.16 ms (Supplementary Table 3; linear mixed effects model; t(58) = 0.69, p = 0.49, 95% confidence interval: -13.17 to 27.49). However, a subject's coefficient for cross-cluster surprisal in the random walk was significantly correlated with that in the 8/29 0100012001400050010001500TrialMean RT (ms)Cross−Cluster EdgeFALSETRUEAPearson's ρ = 0.8p < 0.001−1000100200300400−1000100200300400Surprisal Effect in Stage 1 (ms)Surprisal Effect in Stage 2 (ms)Dn=301357911131513579111315NodePrevious Node10501150RT (ms)Bn=162n=59C5151423 Hamiltonian walk (Fig. 2D; t(57) = 10.089, p < 0.001, Pearson's correlation coefficient r = 0.8, 95% confidence interval: 0.69 to 0.88), suggesting a diminished yet persistent effect of topological edge role after eliminating local repetitions. We verified that this result was not the result of idiosyncrasies in assignment of motor actions to nodes by performing a permutation test where the identity of cross-cluster edges was randomly assigned (see Supplementary Fig. 3). While the cross-cluster surprisal effect is a useful measure of how well a modular graph is being learned, it is not a measure that generalizes to non-modular structures. To quantitatively examine the learnability of graph structure across many graph topologies, it would be useful to develop a generalizable measure of the learnability of single transitions from one button press to another. In developing such a measure, it is important to note that two potential explanations exist for improvement in response to a given target: (i) improvement is node-dependent (for all edges leading to that node), or (ii) improvement is edge-dependent, with the rate of improvement depending on the preceding node. Notably, the inclusion of additional edges (with the same set of nodes) in stage 2 of Experiment 1 led to a large increase in mean reaction time (Fig. 3A). To verify that subjects had learned the edges versus nodes of the graph, we examined stage 2 of Experiment 1: when subjects were shown (and asked to respond to) a sequence of stimuli drawn from a fully connected graph. We labeled each edge as either previously learned or not previously learned, based on the respective motor action assigned to the endpoint nodes and whether the edge (sequence of motor actions) was present in the graph learned by that subject in the first stage. We then estimated learner sensitivity to new edges as measured by a change in reaction time, which we referred to as the novel edge effect (see Methods). Using a mixed effects model, we found that subjects were significantly slower when responding to edges that had not previously been seen, with an expected increase of 25.5 ms (linear mixed effects model; Supplementary Table 4; t(131) = 4.35, p < 0.001, 95% confidence interval: 14.01 to 37.0). This finding supports the notion that subjects learn single edges in a graph. However, altering the number of edges in the graph decreases predictability of transitions from a single node. While subjects were faster on previously seen edges, this difference in reaction time could be attributable to improvements in compound motor movements, rather than to the learning of any higher-order structure. We therefore asked whether reaction time improvements for sequences generated by each graph type might modulate this novel edge effect. In other words, was it the case that subjects with greater sensitivity to graph structure would be more affected by disruptions to it? Further, might this association between learning measures differ by graph type? We estimated each subject's learning rate27, and we also estimated the novel edge effect for each subject. We found that faster learners showed a significantly greater novel edge effect in the second stage of the experiment than slower learners (Fig. 3B; t(107) = 3.79, p < 0.001 Pearson's correlation coefficient r = 0.34, 95% confidence interval: 0.17 to 0.50). When subdivided by graph type (Fig. 3C), this effect was significant for the 9/29 modular graph (t(28) = 2.93, p = 0.007, Pearson's correlation coefficient r = 0.48, 95% confidence interval: 0.15 to 0.72) and the lattice graph (t(41) = 2.28, p = 0.027, Pearson's correlation coefficient r = 0.34, 95% confidence interval: 0.04 to 0.58), but not significant for the random graph (t(34) = 1.31, p = 0.2, Pearson's correlation coefficient r = 0.22, 95% confidence interval: -0.11 to 0.51). Intriguingly, this pattern of results suggests that subjects can more easily learn the regular structure of modular and lattice graphs, and display slower reaction times when expectations are violated. We note, however, that the difference between the modular and lattice conditions and the random condition was not in itself significant (modular and random: Fisher's z = 1.18, one-sided p = 0.12, lattice and random: Fisher's z = 0.55, one-sided p = 0.29.) Next, we tested whether certain graph structures facilitate learning more than others. We predicted that sequences generated by the modular graph would be the easiest for participants to learn, due to the graph's segregated meso-scale structure. As subject groups were exposed to different pairs of graph topologies, we performed three separate within-subject analyses using the data from Experiment 3. Each analysis examined a pair of graph types, with the order of exposure counterbalanced between subjects. For example, the first group was composed of (i) subjects first exposed to a stream of stimuli produced by a random walk on the lattice graph, followed by a stream of stimuli produced by a random walk on the random graph, as well as (ii) subjects first exposed to a stream of stimuli produced by a random walk on the random graph, followed by a stream of stimuli produced by a random walk on the lattice graph. In the same manner, the second group corresponded to modular/lattice, and the third group corresponded to modular/random. We separately fit a mixed effects model to each group. We found that the modular graph elicited significantly quicker responses than both the lattice (linear mixed effects model; t(70) = 2.35, p = 0.022; expected difference of 34.89 ms; 95% confidence interval: -44.49 to -4.02) and random (t(69) = 3.429, p = 0.001; expected difference of -34.89 ms; 95% confidence interval: -54.82 to -14.95) graphs (Fig. 3D,E). We did not find a significant difference between the lattice and random graphs (t(68) = 1.48, p = 0.14; expected difference of 12.85 ms; 95% confidence interval: -29.88 to 4.17). Models are summarized in Supplementary Table 5. These findings support the hypothesis that the presence of meso-scale structure in modular graphs impacts learnability. In a final set of analyses, we investigated the extent of the influence of graph structure on learning. More specifically, we tested whether smaller scale topological features or larger scale topological features might also impact learning, in addition to the meso-scale features studied in the previous section. First, we studied smaller scale topological features using degree, a summary statistic of a node's neighborhood defined by the number of edges emanating from a node. Second, we examined large scale topological features using betweenness centrality, which intuitively captures a node's role in mediating long distance traversals through the graph, and which is defined by the fraction of shortest paths that pass through a given node (Fig. 4A). We studied the 10/29 Figure 3. Learning Rate and Edge Surprisal. Impact of new edges on reaction time. (A) Mean RT increases in stage 2 when new edges are added to the graph (trials 1501-2000). Included for comparison are Experiments 2 and 3, where -- respectively -- only the walk or a subset of edges were changed. In both cases the increase in RT is much smaller. (B) Per-subject learning rate correlated with the novel edge effect, defined as the mean difference in reaction time for a subject learning the second graph when responding to a novel edge versus a familiar edge (see Methods; n=109). Learning rate, the model coefficient for log(trial), was scaled amongst all subjects to the range [0,1]. The blue line is the least squares fit, with the gray envelope indicating the 95% confidence interval. (C) Individual correlations shown for the three types of graphs trained on in the first stage. Subjects exposed to the modular and lattice graphs show a significant relationship (p < 0.01,n = 30 and p < 0.03,n = 43, respectively), while those exposed to the random graph do not (p < 0.2,n = 36). Solid lines represent least squares fits, and gray envelopes represent the respective 95% confidence intervals. (D) Differences in reaction time by graph type, across graphs learned in sequence. Each bar shows the number of milliseconds by which the modeled effect for the top listed graph is faster. The increase in RT from lattice to modular, and from random to modular graphs, are both significant to p = 0.02 and p = 0.001, respectively (See Supplementary Table 5). Error bars indicate standard error as estimated in the mixed effects model. Asterisks indicate significance in the mixed effects model. Group sizes: Lattice-Random: n=70, 11/29 Modular-Lattice: n=72, Modular-Random: n=71. (E) Examples of the graph types. Stage 210001200140010001250150017502000TrialMean RT (ms)Experiment1 (n=109)2 (n=59)3 (n=213)ABC010203040***Significant in Mixed Effects ModelLattice vs.RandomModular vs.LatticeGraph ComparisonModular vs.Random=RT Difference (ms)DLatticeModularRandomE0Stage 10Pearson's ρ = 0.34p < 0.0010501001500.000.250.500.751.00Scaled Learning RateNovel Edge Effect (ms)n=1090501001500.000.250.500.751.00Scaled Learning RateNovel Edge Effect (ms)Stage 1 GraphModular (n=30)Lattice (n=43)Random (n=36)n=70n=72n=71 relation between reaction time and these statistics specifically in the random graph exposures from the first and third experiment, where we observed the greatest variability in degree and betweenness centrality over nodes. In all cases, we regressed out the visits to a node, to separate the influence of network topology from the influence of increased exposure. We found that node degree highly predicted the mean response time on a node (Fig. 4B,D; Kendall's τ = 0.072,n = 2655, p < 0.001), as did node betweenness centrality (Fig. 4C,E; Kendall's τ = 0.044,n = 2655, p < 0.001). These results indicate that not only does meso-scale graph organization affect learnability, but so do smaller scale topological features quantifying the number of edges in a node's immediate neighborhood, and larger scale topological features quantifying a node's role in long-distance traversals through the graph. Intriguingly, we observed an inverted relationship when we refrained from regressing out the number of visits to a node, a fact that highlights the complex relationship between topology and learnability. (For full results, and additional findings related to other graph metrics, see Supplementary Fig. 4-7. Also note that degree and node betweenness centrality were correlated in the graphs analyzed in this experiment -- Kendall's τ = 0.669,n = 2655, p < 0.001 -- indicating that nodes with dense local connectivity also play an important role in long-distance traversals.) In turning to a discussion of our results, we begin by grounding our experimental setup and findings in the context of prior literature. Then in later sections, we turn our attention to a discussion of more specific implications of graph architecture for learning. In this study, we capitalize on a rich history of using motor response times as a learning measure (e.g., Cleeremans & McClelland 11; for reviews see: Cleeremans et al. 28, Robertson 29). While those studies (and much of the artificial grammar learning work since Reber 30) have clear parallels to the present findings, our experiments diverge in the fundamental question they address. While we similarly generate sequences by "walking" along the edges of a graph, we additionally systematically manipulate the topological properties of the graph underlying motor responses. In contrast to the bulk of the artificial grammar learning literature, which largely relies on an arbitrary finite-state grammar, we instead apply organizational principles informed by graph theory to study biases in human learning. In the current study, we employed tools from the field of network science to determine whether and how graph structure influences sequence learning. More specifically, our experiment built upon previous work demonstrating that modular graph organization influences visual statistical learning3,4. We extended this line of inquiry to the motor domain by assigning unique button presses to nodes of a modular graph, generating sequences via a random walk on that graph, and asking whether learners exhibit similar behavioral effects at the boundaries between clusters of nodes. Indeed, we found that learners displayed a sharp increase in reaction time when transitioning from one cluster of button presses to another, indicating that they developed implicit expectations about the underlying topology of incoming sequential input. The sum of these results indicates that 12/29 Figure 4. Relation Between Small and Large Scale Graph Statistics and Reaction Time. (A) Illustration of node betweenness centrality. We show shortest paths from a number of nodes on the far left to a node on the far right, which all pass through the blue node. (B) Relationship between node degree and reaction time (RT), after regressing out the number of visits to a node, with each point representing a separate node, e.g., 15 points per subject. The regression line shows least squares fit, and the gray envelope is the 95% confidence interval. Reported correlation is based on Kendall's τ (n=177 subjects). (C) Relationship between node betweenness centrality and reaction time using the same approach as used with node degree. (D) Mean reaction time shown as a function of degree, where the mean was z-scored across the 15 nodes for a given subject. Error bars represent bootstrapped 95% confidence intervals. (E) Mean reaction time as a function of node betweenness centrality using the same approach as used with node degree. 13/29 τ=0.072p<0.001−0.50.00.51.00369Node DegreeRT (Node Visits Regressed Out)Bτ=0.044p<0.001−0.50.00.51.00.00.10.20.30.4Node Betweenness CentralityRT (Node Visits Regressed Out)C−0.20.00.20369Node DegreeRT (Node Visits Regressed Out)Least Squares FitMeanD−0.20.00.20.00.10.20.30.4Node Betweenness CentralityRT (Node Visits Regressed Out)Least Squares FitMeanEAn=177 subjects learners capitalize on modular structure in developing expertise in executing complex motor sequences. We further generalized our observations to other graph topologies, and showed that certain local-, meso-, and global-scale features of graph architecture are associated with higher learning rates than others, suggesting a critical impact of graph topology on motor skill acquisition. One important and outstanding question is the following: What exactly forms the basis for the observed increase in RT when switching between clusters in the modular graph? One could imagine a scenario in which the importance of the topological role played by the cross-cluster edges would lead to an optimization of the associated motor movements, and a decrease (rather than an increase) in RT. Our expectation of an increase in RT was primarily based on prior observations in similar tasks, particularly the visual perception task reported in Karuza et al. 4. The fact that we observe the same increase in RT at cross-cluster edges in a motor task suggests that the surprisal effect is a general (rather than modality-specific) property of probabilistic sequential learning of modular structures. Yet, its presence across task modalities does not equate to a cause. Here we describe several possible mechanisms for the surprisal effect at a number of different explanatory levels. One level of explanation begins with the underlying neural processes. Using a visual item rotation detection task based on a graph traversal, Schapiro et al. 3 measured the distinguishability of clusters by asking subjects to mark natural 'segmentation' points. By studying fMRI data acquired during the performance of the task, the authors found that patterns of BOLD activity were more similar for items within clusters than for items between clusters, despite the fact that visual stimuli were randomly assigned to nodes. Based on this work, it is natural to ask whether a similar mechanism might apply to the processing of targets for our motor response task. If so, then similar neural representations of nodes within a cluster might allow for quicker responses to targets that are nearby in representational similarity space. It would be interesting to test this possibility in a future study that combined neuroimaging with the behavioral assay we provide in our study. Neural processes aside, we have referred to the increase in RT at cross-cluster edges as a surprisal effect based on prior work4. Nevertheless, alternative interpretations -- beyond surprise -- also exist. For instance, the increase in RT when entering a new cluster might be construed as analogous to the sequence initiation cost commonly observed in SRT8 and DSP31 tasks, where increases in RT are observed at the beginning of separable chunks of responses. The idea that our observed surprisal effect may in fact represent a preparatory cost is certainly plausible. However, it remains an open question how a sequence initiation cost might generalize to a situation such as the current task in which subsequences are less structured. Indeed, it would be interesting in the future to assess evidence for a type of sequence initiation cost that reflects preparation for a regime of likely (rather than fixed) responses. The surprisal effect also shows strong similarities to the so-called switch cost observed in the task switching literature32. 14/29 In common parlance, the switch cost is an increase in response time when learners are prompted to switch tasks between trials, where a task might be reporting the color of a stimulus, or the magnitude of a number. Typically each task is conceived of as having its own stimulus-response mapping, which is similar to the discrete set of responses anticipated within each of our clusters. While many studies have demonstrated the persistence of the switch cost even when subjects anticipate task order, a few studies have importantly shown that a switch cost is still observed when participants are led to implicitly learn a task sequence33,34, similar to the implicit presentation of modular structure in our current study. Moreover, the relationship between task and sequence has been shown to be hierarchical35, where both task and sequence interact with one another. This observation provides a link to our current work where not only do the two interact, but the task (in the form of a cluster) is defined by the sequence. In many real-world situations, tasks do not have explicitly defined subtasks nor clear boundaries between those subtasks, and natural divisions are only learned through experience. Thus, a comprehensive exploration of the flexible interplay between task and task sequence might be a promising direction for future research. One final framing of interest relates to the dependence of RT on the local topology as measured by the degree and betweenness centrality of the random graph. Both of these metrics capture the distinctiveness of a node's role within the graph. The degree reflects the number of accessible nodes from a single node; the betweenness centrality reflects the likelihood that a node will be necessary to traverse when moving between any pair of nodes in the graph. Importantly, the modular graph also has important local structure, and the nodes connecting two clusters in the modular graph (boundary nodes) serve a distinct role from the nodes existing within a cluster. One possibility is that a similar mechanism underlies the modulation of RT by the cluster structure and the modulation of RT by the degree and betweenness centrality structure. For instance, boundary nodes in the modular graph exhibit much higher betweenness centrality (0.22) than within-cluster nodes (0.04). And indeed, in both the random graphs and the modular graphs, higher betweenness centrality is significantly associated with higher RTs when differences in degree are accounted for. Thus one could conceive of the surprisal effect as a direct result of these graph properties. However, our current experiment does not allow us to further explore this relationship, as subjects only learned a single modular graph with two distinct classes of nodes. A potential future direction would be to explore whether this is a general relationship across other graphs with modular structure. The primary aim of this study was to examine differences in learners' sensitivity to distinct graph structures, while holding constant the process through which these graphs were traversed (i.e., via a random walk). By exposing participants to modular, lattice, and random graphs, we sought to ascertain how higher-level structure might aid or impede motor skill acquisition. Given that modularity is an essential organizational principle underlying such varied systems as music36 and social networks21,37, 15/29 we anticipated a privileged role for this form of information structure, relative even to sequences generated from the highly structured lattice graphs. In an initial between-subjects experimental design, we provided evidence that learners tracked pairwise statistics, or edges linking nodes across all three graph structures. However, the extent to which learners displayed sensitivity to novel edges was predicted by learning rate only in the lattice and modular graphs. In other words, the overall timecourse of learning throughout exposure to sequences generated by a random graph was not associated with sensitivity to local transition statistics on subsequent measures. We therefore propose that graphs featuring regular structural organization (i.e., lattice and modular) might serve to boost knowledge of local regularities. Further, by capitalizing on a complementary within-subjects experimental design, we directly contrasted the acquisition of sequences generated by distinct graph structures. Compellingly, learning rates for the modular graph condition were significantly faster relative to both the lattice and random graph conditions. While the differences between the highly structured modular condition and the relatively unstructured random condition were perhaps to be expected, the differences between the modular condition and the lattice condition were uniquely insightful. In particular, nodes within the lattice and modular graphs were precisely equated in degree. They only differed in their meso-scale architecture, wherein neighboring nodes within the modular graph were densely interconnected with one another. While the lattice graph was also highly structured, it lacked this key organizational property, to the detriment of the learner, which demonstrates that the learned pairwise associations do not capture the full scope of learners' pattern sensitivity. Instead, we provide evidence that learners clearly benefit from modularity when it underpins the generation of complex motor sequences. Notably, our performance measure associated with modularity, the surprisal effect, persisted even when altering graph topology and the walk taken upon that topology. However, the effect observed when we considered an altered transition structure was significantly weaker than the effect observed when we considered the original transition structure. We believe that we have ruled out simple confounds, particularly in having shown that the relationship between reaction time on the random and Hamiltonian walks is specific to those edges that bridge clusters, and not an artifact of our analysis methods. However, ruling out these simple confounds is not wholly satisfying, and it remains an open question whether the weakening of the surprisal effect reflects limited training in the first stage of the experiment, or whether learners discard their previous response biases as they adapt to the new statistical structure of the second stage of the experiment. Regarding the first point, while sensitivity to statistical structure emerges in a short time frame, persistence and generalization to new contexts may require more extensive training. Likewise, it would interesting to retest our current experimental setup, but without a shift to the Hamiltonian walk, to ask whether the break itself disrupts previously learned statistics. We note that learned statistics are particularly sensitive to contextual shifts38, and therefore it is possible 16/29 that the division between the two sections of the experiment was too explicit given the short timeframe. Fully addressing this possibility will require further data collection in future. The graph-specific effects that we observed indicated that meso-scale graph architecture impacts learning. Yet, these data do not address whether meso-scale architecture alone is privileged, or whether both smaller and larger scale topological features also play a role in the learning process. Using the random graph exposures, which had the greatest variability in multiscale network architecture, we studied (i) a measure of small scale topological structure in the node degree, which captures the extent to which a given node is connected to other nodes in a network, and (ii) a measure of large scale topological structure in the node betweenness centrality, which incorporates information about the role of a node within the entire graph by measuring its importance in shortest paths between other node pairs. We found that nodes of higher degree and betweenness centrality were associated with lower reaction times, as might be expected simply from the fact that learners would be more frequently exposed to these nodes via a random walk through the graph. Unexpectedly, however, after regressing out the number of times each node was visited, these relationships were inverted such that nodes of higher degree and betweenness centrality were associated with higher reaction times. This finding has important implications for how we understand the impact of smaller and larger scale architecture on learning. Specifically, at small topological scales, we propose that when learners are exposed to sequences generated by a heterogeneous topology such as is present in random graphs, a trade-off exists between repeated exposure to a given node (i.e., due to high degree) leading to a lower reaction time, and the complex representation introduced by the high number of its neighbors leading to a higher reaction time. At large topological scales, we similarly propose that a trade-off exists between repeated exposure to a given node due to its location along shortest paths in the graph, and the complexity of the possible paths along which it could be visited. In our current study, degree and betweenness centrality were significantly correlated with stimulus exposure. These nodes thus displayed a tradeoff between (i) familiarity, as modulated by the frequency of exposure, and (ii) uncertainty, as modulated by the probability of the future state being narrowly versus widely spread amongst motor actions. Understanding this potential tradeoff remains an important area for future work. One tractable strategy could be to consider other, non-random walks on the graph that would allow a node to exhibit both low familiarity and high uncertainty, or vice versa. It could also be useful to consider graph topologies that would allow variation in betweenness centrality without impacting degree, so as to disambiguate the effect of one versus the other. For example, graphs could be constructed in which low-degree nodes serve as bridges between disconnected areas of the graph, thus having high betweenness centrality. Separately modulating different local statistics on the graph, as well as the type of walk used, could allow for an expanded understanding of the topological drivers of the observed 17/29 RT variation. The separability of perceptual and motor learning in this and similar tasks continues to be a matter of debate. For example, Deroost & Soetens 39 found that, in the context of an SRT task, stimulus-stimulus (i.e., perceptual) learning was limited to simple deterministic sequences. After training participants on more complex probabilistic sequences, they did not find evidence of perceptual learning. The separability of perceptual and motor learning systems is a challenging issue and one that is not yet fully resolved (e.g., it is likely that one system bolsters the other). Unfortunately, because the present set of experiments was not designed to address this issue, we cannot make strong claims one way or the other. Thus, we elect to maintain the most conservative interpretation of the data possible: our task involves sequences of motor responses, so we frame our results as evidence of motor skill learning. However, teasing apart perceptual and motor learning under this framework is a fascinating area for future study. Especially when considering closely related prior work3,4, we suggest that our observed pattern of results is likely reflective of domain-general processes. We note a few methodological considerations that are particularly pertinent to this work. First, there exists a broad literature on the theory of graph structure as well as on structures found in natural stimuli. Here we sample only a small portion of possible graphs representing stimulus relationships. We study two stereotyped graphs, one with meso-scale clustering (modular) and one with no meso-scale clustering (lattice). However, graphs can exhibit other diverse topologies such as core-periphery structure, as well as other configurations of both high and low clustering beyond those tested. Moreover, the graphs we examine are all comprised of 15 nodes. While impractical for the current study, the relationships between natural stimuli might best be represented by graphs composed of hundreds or even thousands of nodes. Thus an open question is how these results generalize to both larger and more diverse networks. Second, we collected no personal information on Mechanical Turk participants. While we screened for eligibility using location and browser, we collected no information on handedness, age, or prior typing experience. However, as our regression models all incorporate per-subject baseline and learning rate effects, we expect minimal impact on our results from any between-subject differences. Third, there exist several limitations to using participants from Mechanical Turk, who might each be viewing the experiment on different browsers and with different levels of accuracy and speed in their internet connection. Fourth, it is possible that reaction time differences might be driven by recency priming. We know that processing times are reduced for a stimulus recently viewed by the learner, and that different nodes within a graph may be differentially affected. This is primarily a concern for modeling the cross-cluster surprisal effect, where transition nodes may have not been seen as recently as pre-transition nodes. However, Karuza et al. 4 found that low level perceptual priming effects did not fully account for the observed cross-cluster surprisal effect. Moreover, all between-graph comparisons are 18/29 dependent on reaction time across the entire graph rather than between different classes of nodes. We also note that excluding priming effects may be overly conservative, given that they may serve as a local cue to graph organization. Fifth and finally, our study does not address the question of how the exploration of graph structures in real stimuli is instantiated in the brain. For future work, it would interesting to consider recent evidence that co-occurrence of visual stimuli leads to increasingly similar neural representations in particular areas of human neocortex40,41, and that these same areas can encode associative distances between objects3,42. These data suggest that an understanding of the sensitivity to the topological properties of graph structures may have important implications in future for an understanding of neural coding. In conclusion, we note that our results highlight the importance of topological structure in learning from a complex environment. We first demonstrate that learning of higher-level statistics operates in the context of an SRT task. Related previous paradigms instead focused on perceptual learning tasks without this complex motor component. Thus, our results suggest that graph-based statistical learning mechanisms are unlikely to be modality-specific. Second, we have examined the impact of systematic differences in graph organization on learning. We find significant differences in subject performance on different graph types, despite identical numbers of stimuli and possible transitions in each graph. In particular, subjects show overall faster reaction times on sequences drawn from the modular graph. Lastly, we begin to explore why sequences drawn from these graphs may be easier or harder to learn, based on node-level statistics that topologically classify the node in relation to both its immediate neighbors and its global role within the graph. Understanding the degree to which these results generalize to other graph structures remains an important direction for future research, as well as understanding whether these results can provide insight on the organization of naturally occurring complex systems. Additionally, our current study does not address the impact of long-term learning on the surprisal effect. Whether the increase in RT persists or disappears as learners become better trained on the task may help distinguish between competing causes. Methods Participants All participants provided informed consent as specified by the Institutional Review Board (IRB) of the University of Penn- sylvania, and study methods and experimental protocols were approved by the IRB. We recruited 381 unique participants to complete our study on Amazon's Mechanical Turk, an online marketplace for crowdsourced work. Worker IDs were used to exclude any duplicate participants, both within and between the three experiments. The entire sample included 109 participants for the first experiment, 59 participants for the second experiment, and 213 participants for the third experiment. No statistical 19/29 methods were used to pre-determine sample sizes but our sample sizes are similar to those reported in previous publications3,4 All participants were financially remunerated for their time. In the first experiment, participants were paid up to $7 for an estimated 40 minutes: $2 for completing each of the two stages, $1 for completing the entire task, and an extra $1 on each stage on which they correctly answered at least 90% of trials. Experiment 2 provided the same payment as Experiment 1, but with an estimated duration of 40 minutes. In the third experiment, subjects were paid up to $11 for an estimated 60 minutes: $3 per stage, $1 for completing the entire task, and $2 for >90% performance on each stage. Experimental Setup Subjects performed a self-paced SRT motor response task using a keyboard. Stimuli were represented as a horizontal row of five gray squares; all five squares were shown at all times during the main phase of the experiment. To indicate a target key or pair of keys that the subject was meant to press, the corresponding squares would be outlined in red (Fig. 1A). When subjects pressed the correct key combination, the squares on the screen would immediately display the next target. If an incorrect key was pressed, or a key was left out of a two-key combination, the message "Error!" was displayed on the screen below the stimuli, and remained until the subject pressed the correct key(s). The squares corresponded spatially with keys 'Space', 'H', 'J', 'K' and 'L', such that the left square represented 'Space' and the right square represented 'L'. These keys were chosen to ergonomically rest underneath the subject's right hand on a QWERTY keyboard with their thumb above 'Space', index finger above 'H', and so on (Fig. 1B). The order in which stimuli were presented to the subject in Experiments 1 and 3 was prescribed by a random walk on a graph of N = 15 nodes connected by E = 30 edges. In each graph, one of the 15 one- or two-finger key combinations was randomly assigned to each node (Fig. 1A). A different graph (mapping of key presses to nodes) was generated for each random walk. For any two nodes that were not connected by an edge, the transition probability was equal to zero. For any two nodes that were connected by an edge, the transition probability was equal to 1 divided by the number of edges emanating from the pre-transition node. In Experiment 2, stage 1 consisted of a random walk as previously described. The order of stimulus presentation in stage 2 was generated by a series of Hamiltonian cycles through the graph. A single cycle consisted of every node in the graph being visited exactly once, and each cycle was followed by another Hamiltonian cycle beginning from a node adjacent from where the last cycle ended. We studied the learning of 3 different graph topologies: a modular graph, a lattice graph, and a random graph (Fig. 1C). The modular graph was characterized by 3 clusters of 5-nodes each, and a greater number of edges between nodes in a cluster than 20/29 between nodes in different clusters. Importantly, each node in the graph had exactly 4 edges, or a degree of k = 4. Thus, for any two nodes that were connected by an edge, the transition probability was equal to 25%. The lattice graph was similar to a ring lattice, in which nodes near one another on the ring tended to be connected to one another. As with the modular graph, each node had exactly 4 edges, or a degree k = 4, thereby creating a flat transition probability of 25% between any two connected nodes. Random graphs were selected out of a possible pool formed by creating 1500 instantiations of the Erdos -- Rényi graph model, all of which we guaranteed were fully connected and had radius of at least 3, meaning there was at least one pair of nodes with a shortest path involving three edges. We then sorted the ensemble by their estimated modularity43, and we discarded graphs with the highest and lowest 2.5% of modularity values. The same pool of random graphs was used across all subjects within a given experiment, though the pool differed between the experiments. In these random graphs, the degree of each node varied from k = 1 to k = 9, and the transition probabilities varied accordingly. Finally, in the second stage of the first experiment, we used a fully connected graph as a point of comparison, in which any node can transition to any other node. In this case, the degree of each node was k = 14, and the probability of transitioning between any two nodes was 1/14 or approximately 7.14%. We ran three separate experiments (see Supplementary Table 1). Each one consisted of two stages that differed in which graph was used to generate the stimulus sequence. The first experiment examined learning as the underlying structure transitioned from a regular graph to a fully connected graph, and did so by comparing reaction times in response to novel versus previously learned edges. The first stage of the first experiment used either a modular graph (n=30), a lattice graph (n=43), or a random graph (n=36) to generate a 1500-node random stimulus walk, while the second stage used a fully connected graph to generate a 500-node walk (random stimulus order). In the second experiment, both stages consisted of a walk over the modular graph. However, the first stage was a 1500-node random walk, while the second stage was a 500-node Hamiltonian walk (n=59). The third experiment employed a within-subjects design to directly compare learning effects between the graph types, accounting for individual variability in baseline reaction times and learning rates. Similar to the first experiment, the first stage of the second experiment consisted of a 1500-node random stimulus walk from either a modular graph, a lattice graph, or a random graph; however, unlike the first experiment, the second stage was a 1500-node random stimulus walk on one of the remaining two graph types. For example, if a subject was shown a sequence of stimuli produced from the modular graph in the first stage, then in the second stage the subject would be shown a sequence of stimuli produced from either the lattice graph or the random graph. For each of the six possible pairs of graphs we collected reaction time data from at least 30 subjects (modular/lattice: n=36, modular/random: n=37, lattice/modular: n=36, lattice/random: n=38, random/modular: n=34, and 21/29 random/lattice: n=32). In all cases, subjects were randomly assigned to experimental conditions. Assignment was done in the experiment code, blinding experimenters to the condition assignment for each individual participant. Subjects were only excluded if they failed to complete the study. Experimental Procedures At the beginning of the first experiment, subjects were provided with the following instructions: "In a few minutes, you will see five squares shown on the screen, which will light up as the experiment progresses. These squares correspond with keys on your keyboard, and your job is to watch the squares and press the corresponding key when that square lights up. This part will take around 30 minutes, followed by a similar task which will take about 5 minutes." To incentivize accuracy on the task, subjects were informed that if they answered more than 90% of trials correctly, they would receive a $2 bonus. Subjects were also instructed "The amount of time the segments take is not fixed, but the number of responses you have to make is. Therefore, you should make your responses both quickly and accurately." While the reward was solely based on accuracy, workers had an implicit incentive to finish quickly due to the fixed reward, allowing more time for other tasks on Mechanical Turk. Before the full experiment began, subjects were given a short quiz to verify that they had read and understood the instructions. If any questions were answered incorrectly, subjects were shown the instructions again and asked to repeat the quiz until they answered all questions correctly. Next, all subjects were shown a 10-trial segment that did not count towards their performance; this segment also displayed text on the screen explicitly telling the subject what keys to press on their keyboard. Subjects then began the 1500-trial stage. A brief reminder was presented before the second stage, but no new instructions were given. After completing the second stage, subjects were presented with performance information and their bonus earned, as well as the option to provide feedback. The second experiment used the same instructions as the first experiment, though the estimated time for the second task was changed to 10 minutes. The third experiment was nearly identical, except that the initial text was changed to reflect the second structured walk: "In a few minutes, you will see five squares shown on the screen, which will light up as the experiment progresses. These squares correspond with keys on your keyboard, and your job is to watch the squares and press the corresponding key when that square lights up. This part will take around 30 minutes, followed by a similar task which will take another 30 minutes." This phrasing in the instructions ensured that learners would differentiate between stages of the experiment, reducing the potential of carry-over effects between learning of graph structures while still preserving the benefit of a within-subject comparison. Both the quiz and 10-trial practice session were still present, and stimulus presentation was identical to that used in the first 22/29 experiment. Mixed effects models were fit using the lme4 library in R (R v3.5.0, lme4 v1.1-17), using the lmer function. Predictors were centered to reduce multicollinearity. All contrasts were orthogonally coded. The observed correlation between fixed effects was less than 0.7. Random effects were chosen as the maximal structure that allowed model convergence, as specified in the next section. All tests were two-sided unless otherwise noted. Analytical Approach For every trial, we computed the reaction time based on the elapsed time from the last button press. We only excluded trials for two reasons: subjects answered incorrectly on their first attempt or the reaction time was implausible (under 100ms or over 5000ms, or more than 3 SDs from their mean reaction time). Effect of Targets on Reaction Time: Since performance was measured based on a key press, it was important to determine whether biomechanical factors related to the use of different fingers, or to different combinations of finger pairs, influenced reaction time. We predicted that one-finger responses would show shorter reaction times than two-finger responses, and that response time would vary based on the finger needed in the response. We calculated the average reaction time for each key press or combination of key presses across subjects and training stages (Fig. 1C). While we observed complex differences in reaction time by finger, we also found robust differences in reaction time driven by the number of fingers required, with one-key presses displaying a shorter reaction time than two-key presses (paired two-tailed t-test for one- and two-finger means for each subject: t(321) = 35.56, p < 0.001; mean difference: 228.86 ms, 95% confidence interval: 216.20 to 241.52 ms). We observed that the relative ordering of keys or key combinations by reaction time was remarkably well-preserved across subjects, and the difference was independent of which graph was seen in the first stage of the experiment (random, lattice, or modular). This observation suggests unique motor timing associated with each target (Supplementary Fig. 8). To ensure that our findings were not systematically biased by differences in reaction time across key presses, we included the target key press as a regressor in all statistical models. Learning Rate: Based on a subject's reaction time profile across the session, we estimated each subject's learning rate from a linear mixed effects model. The learning rate estimate was the per-subject random effect for trial number. A faster learning rate was captured as a negative coefficient, which indicated that a subject's reaction time decreased more rapidly over time. The model was fit to the first stage of the first experiment, with formula RT ∼ log(Trial)∗Graph+Target+(1+log(Trial) Subject), where Target represented the key combination of the target node, controlling for biomechanical differences in motor response, 23/29 Graph was one of random/lattice/modular, and Trial was the sequential trial number (1 to 1500). We verified experimentally that log(Trial) provided a substantially better fit than Trial (See Supplementary Fig. 9). The log transformation served to increase the normality of the data, although formal testing of the degree of normality was not performed. Surprisal Effect: One measure reflective of meso-scale graph structure is the surprisal effect, defined as an increase in reaction time when transitioning to a new cluster as compared to any previous within-cluster reaction times4. To measure this effect in our task, we fit a linear mixed effects model of the form RT ∼ log(Trial)∗ EdgeType + Target + (1 + log(Trial)∗ EdgeType Subject) to data acquired during the first stage of the experiment on the modular graph where RT is reaction time, and where EdgeType indicated whether an edge was within or between clusters. From this model, we examined the model coefficient for EdgeType. Surprisal Transfer: We fit the surprisal effect model separately to the stage 1 (random walk) and stage 2 (Hamiltonian walk) data from Experiment 2. We estimated the per-subject effect of EdgeType as a measure of surprisal, and additionally examined the Pearson correlation coefficient between each subject's surprisal effect in stage 1 versus surprisal effect in stage 2. Novel Edge Effect: Next, we sought to examine whether subjects were specifically improving their performance at transitions present in the graph and therefore sensitive to violations of the learned structure of the graph. We defined a LearnedEdge variable for all edges in stage 2 that was true if an edge between the set of finger combinations had been present in the first stage, and false otherwise. We then computed a novel edge effect measure as the coefficient for learned versus unlearned edges in a linear mixed effects model fit to the fully connected graph data using RT ∼ log(Trial)∗ Graph∗ LearnedEdge + Target + (1 + log(Trial)∗ LearnedEdge Subject). Graph Effects: In the third experiment, we investigated whether differences in meso-scale structure affected learnability by quantifying reaction time differences due to graph type. Since not all subjects were exposed to all graph types, we split the data into three groups based on exposure: modular/lattice, random/modular, and random/lattice, with each group roughly split by which graph was used for the first stage versus for the second stage. We fit a linear mixed effects model to each of the three data subsets, in order to estimate whether graph type was a significant effect in each model: RT ∼ log(Trial)∗ Graph∗ Stage + Target + (1 + log(Trial)∗ GraphSubject), where Trial varied between 1 and 1500, Graph was the current graph of the two graphs that the subject saw, and Stage corresponded to either the first stage or the second stage. Node Effects: After determining the effect of graph type, we finally turned to quantifying the impact of node-level statistics, or those that could explain reaction time differences within a single graph. We examined several traditional graph metrics: degree, clustering coefficient, node betweenness centrality, and edge betweenness centrality. The degree is defined as the 24/29 number of edges connecting to a given node, given by ki = ∑ j Ai j where A is the adjacency matrix. The clustering coefficient can be defined as the fraction of possible edges between a node's neighbors, given by Ci = 2Li ki(ki−1) where Li is the number of edges between any two neighbors of node i. The node betweenness centrality is defined as the fraction of shortest paths in the graph that pass through node v, given by CB(v) = ∑s(cid:54)=v(cid:54)=t σst (v) σst where σst (v) is the number of shortest paths from node s to node t that pass through node v, and σst is the total number of shortest paths in the graph from node s to node t. The edge betweenness centrality is defined as the fraction of shortest paths passing through an edge e, given by CB(e) = ∑s(cid:54)=e(cid:54)=t σst (e) σst , where now e refers to an edge rather than a node. Data Availability The data that support the findings of this study are available from the corresponding author upon reasonable request. Code Availability The code that supports the findings of this study is available from the corresponding author upon reasonable request. 25/29 References 1. Aslin, R. N. & Newport, E. L. Statistical Learning: From Acquiring Specific Items to Forming General Rules. Current Directions in Psychological Science 21, 170 -- 176. ISSN: 0963-7214 (2012). 2. Newman, M. E. J. Networks: An Introduction (Oxford University Press, 2010). 3. Schapiro, A. C., Rogers, T. T., Cordova, N. I., Turk-Browne, N. B. & Botvinick, M. M. Neural representations of events arise from temporal community structure. Nature Neuroscience 16, 486 -- 492. ISSN: 1097-6256 (Feb. 2013). 4. Karuza, E. A., Kahn, A. E., Thompson-Schill, S. L. & Bassett, D. S. Process Reveals Structure: How a Network Is Traversed Mediates Expectations About Its Architecture. Scientific Reports 7, 12733 (2017). 5. Newman, M. E. J. Complex Systems: A Survey. Am. J. Phys. 79, 800 -- 810 (2011). 6. Saffran, J. R., Aslin, R. N. & Newport, E. L. Statistical Learning by 8-Month-Old Infants. Science 274, 1926 -- 1928. ISSN: 0036-8075 (1996). 7. Nissen, M. J. & Bullemer, P. Attentional requirements of learning: Evidence from performance measures. Cognitive Psychology 19, 1 -- 32. ISSN: 00100285 (Jan. 1987). 8. Hunt, R. H. & Aslin, R. N. Statistical learning in a serial reaction time task: Access to seperable statistical cues by individual learners. Journal of experimental psychology. General 130, 658 -- 680. ISSN: 0096-3445 (2001). 9. Fiser, J. & Aslin, R. N. Statistical learning of higher-order temporal structure from visual shape sequences. Journal of Experimental Psychology: Learning, Memory, and Cognition 28, 458 -- 467. ISSN: 0278-7393 (2002). 10. Turk-Browne, N. B., Jungé, J. A. & Scholl, B. J. The Automaticity of Visual Statistical Learning. Journal of Experimental Psychology: General 134, 552 -- 564. ISSN: 1939-2222 (2005). 11. Cleeremans, A. & McClelland, J. L. Learning the structure of event sequences. Journal of Experimental Psychology: General 120, 235 -- 253. ISSN: 1939-2222 (1991). 12. Furl, N. et al. Neural prediction of higher-order auditory sequence statistics. NeuroImage 54, 2267 -- 2277. ISSN: 10538119 (Feb. 2011). 13. Newport, E. L. & Aslin, R. N. Learning at a distance I. Statistical learning of non-adjacent dependencies. Cognitive Psychology 48, 127 -- 162. ISSN: 00100285 (2004). 26/29 14. Gómez, R. L. Variability and Detection of Invariant Structure. Psychological Science 13, 431 -- 436. ISSN: 0956-7976 (Sept. 2002). 15. Bollobas, B. Random Graphs (Cambridge University Press, 2001). 16. Jarvis, J. P. & Shier, D. R. in Applied mathematical modeling: a multidisciplinary approach (1999). doi:10.1.1.39. 1937. 17. Goldstein, R. & Vitevitch, M. S. The influence of clustering coefficient on word-learning: how groups of similar sounding words facilitate acquisition. Frontiers in Psychology 5, 2009 -- 2014. ISSN: 1664-1078 (Nov. 2014). 18. Bales, M. E. & Johnson, S. B. Graph theoretic modeling of large-scale semantic networks. J Biomed Inform 39, 451 -- 464 (2006). 19. Vitevitch, M. S. What can graph theory tell us about word learning and lexical retrieval? J Speech Lang Hear Res 51, 408 -- 422 (2008). 20. Palla, G., Barabasi, A. L. & Vicsek, T. Quantifying social group evolution. Nature 446, 664 -- 667 (2007). 21. Girvan, M. & Newman, M. E. J. Community structure in social and biological networks. Proceedings of the National Academy of Sciences 99, 7821 -- 7826. ISSN: 0027-8424 (June 2002). 22. Karuza, E. A., Thompson-Schill, S. L. & Bassett, D. S. Local Patterns to Global Architectures: Influences of Network Topology on Human Learning. Trends Cogn Sci 20, 629 -- 640 (2016). 23. Steyvers, M. & Tenenbaum, J. B. The large-scale structure of semantic networks: Statistical analyses and a model of semantic growth. Cognitive Science 29, 41 -- 78. ISSN: 03640213 (2005). 24. Mengistu, H., Huizinga, J., Mouret, J. B. & Clune, J. The Evolutionary Origins of Hierarchy. PLoS Comput Biol 12, e1004829 (2016). 25. Hermundstad, A. M. et al. Variance predicts salience in central sensory processing. eLife 3, 1 -- 28. ISSN: 2050084X (2014). 26. Heathcote, A., Brown, S. & Mewhort, D. J. The power law repealed: the case for an exponential law of practice. Psychonomic bulletin & review 7, 185 -- 207 (June 2000). 27. Karuza, E. A., Farmer, T. A., Fine, A. B., Smith, F. X. & Jaeger, T. F. On-line measures of prediction in a self-paced statistical learning task. Proceedings of the 36th Annual Meeting of the Cognitive Science Society (2014). 27/29 28. Cleeremans, A., Destrebecqz, A. & Boyer, M. Implicit Learning: News From the Front. Trends in Cognitive Sciences 2, 406 -- 416 (1998). 29. Robertson, E. M. The Serial Reaction Time Task: Implicit Motor Skill Learning? Journal of Neuroscience 27, 10073 -- 10075 (2007). 30. Reber, A. S. Implicit Learning of Artificial Grammars. Journal of Verbal Learning and Verbal Behavior 6, 855 -- 863 (1967). 31. Verwey, W. B., Abrahamse, E. L. & de Kleine, E. Cognitive Processing in New and Practiced Discrete Keying Sequences. Frontiers in Psychology 1, 1 -- 13 (2010). 32. Kiesel, A. et al. Control and Interference in Task Switching-A Review. Psychological Bulletin 136, 849 -- 874 (2010). 33. Koch, I. Automatic and Intentional Activation of Task Sets. Journal of Experimental Psychology: Learning, Memory, and Cognition 27, 1474 -- 1486 (2001). 34. Gotler, A., Meiran, N. & Tzelgov, J. Nonintentional Task Set Activation: Evidence From Implicit Task Sequence Learning. Psychonomic Bulletin & Review 10, 890 -- 896 (2003). 35. Schneider, D. W. & Logan, G. D. Hierarchical control of cognitive processes: Switching tasks in sequences. Journal of Experimental Psychology 135, 623 -- 640 (4 2006). 36. Gleiser, P. & Danon, L. Community structure in jazz. Adv. Compl. Syst. 6, 565 -- 573 (2003). 37. Tompson, S. H., Kahn, A. E., Falk, E. B., Vettel, J. M. & Bassett, D. S. Individual differences in learning social and nonsocial network structures. Journal of Experimental Psychology: Learning, Memory, and Cognition. ISSN: 1939-1285 (July 2018). 38. Gebhart, A. L., Aslin, R. N. & Newport, E. L. Changing structures in midstream: Learning along the statistical garden path. Cognitive Science 33, 1087 -- 1116. ISSN: 03640213 (2009). 39. Deroost, N. & Soetens, E. Perceptual or motor learning in SRT tasks with complex sequence structures. Psychological Research 70, 88 -- 102 (2006). 40. Messinger, A., Squire, L. R., Zola, S. M. & Albright, T. D. Neuronal representations of stimulus associations develop in the temporal lobe during learning. Proceedings of the National Academy of Sciences 98, 12239 -- 12244. ISSN: 0027-8424 (Oct. 2001). 28/29 41. Li, N. & DiCarlo, J. J. Unsupervised Natural Experience Rapidly Alters Invariant Object Representation in Visual Cortex. Science 321, 1502 -- 1507. ISSN: 0036-8075 (2008). 42. Garvert, M. M., Dolan, R. J. & Behrens, T. E. A map of abstract relational knowledge in the human hippocampal -- entorhinal cortex. eLife 6, 1 -- 20. ISSN: 2050-084X (Apr. 2017). 43. Newman, M. E. Modularity and community structure in networks. Proc Natl Acad Sci U S A 103, 8577 -- 8582 (2006). Acknowledgments We thank David M. Lydon-Staley and Steven H. Tompson for feedback on earlier versions of this manuscript. This work was supported by the National Science Foundation CAREER award to DSB (PHY-1554488), the Army Research Laboratory through contract number W911NF-10-2-0022, and the Army Research Office through contract number Grafton-W911NF-16- 1-0474 and contract number DCIST- W911NF-17-2-0181. We also acknowledge additional support from the John D. and Catherine T. MacArthur Foundation, the Alfred P. Sloan Foundation, the ISI Foundation, the Paul Allen Foundation, the Army Research Office (Bassett-W911NF-14-1-0679), the Office of Naval Research, the National Institute of Mental Health (2-R01-DC-009209-11, R01 MH112847, R01-MH107235, R21 MH-106799, R01 MH113550), the National Institute of Child Health and Human Development (1R01HD086888-01), the National Institute of Neurological Disorders and Stroke (R01 NS099348), and the National Science Foundation (BCS-1441502, BCS-1631550, and CNS-1626008). The content is solely the responsibility of the authors and does not necessarily represent the official views of any of the funding agencies. The funders had no role in study design, data collection and analysis, decision to publish or preparation of the manuscript. Competing Interests The authors declare no competing interests. Author Contributions A.E.K., E.A.K., J.M.V., and D.S.B. conceived the project. A.E.K., E.A.K., J.M.V., and D.S.B. planned the experiments and analyses. A.E.K. performed the experiments and analyses. A.E.K., E.A.K., and D.S.B. wrote the manuscript and Supplementary Information. E.A.K., J.M.V., and D.S.B. edited the manuscript and Supplementary Information. 29/29
1210.2140
1
1210
2012-10-08T03:41:45
Natural electromagnetic waveguide structures based on myelin sheath in the neural system
[ "q-bio.NC", "cond-mat.soft", "physics.bio-ph" ]
The saltatory propagation of action potentials on myelinated axons is conventionally explained by the mechanism employing local circuit ionic current flows between nodes of Ranvier. Under this framework, the myelin sheath with up to 100 layers of membrane only serves as the insulating shell. The speed of action potentials is measured to be as fast as 100 m/s on myelinated axons, but ions move in fluids at just 100 nm/s in a 1 V/m electric field. We show here the action potentials, in the form of electromagnetic (EM) pulses, can propagate in natural EM waveguide structures formed by the myelin sheath merged in fluids. The propagation time is mainly cost on the duration for triggering EM pulses at nodes of Ranvier. The result clearly reveals the evolution of axons from the unmyelinated to the myelinated, which has remarkably enhanced the propagation efficiency by increasing the thickness of myelin sheath.
q-bio.NC
q-bio
Natural electromagnetic waveguide structures based on myelin sheath in the neural system Jiongwei Xue and Shengyong Xu Department of Electronics, and Key Laboratory for the Physics & Chemistry of Nanodevices, School of Electronics Engineering and Computer Science, Peking University, Beijing 100871, People’s Republic of China. Jiongwei Xue, [email protected] Shengyong Xu, [email protected] The saltatory propagation of action potentials on myelinated axons is conventionally explained by the mechanism employing “local circuit ionic current flows” between nodes of Ranvier. Under this framework, the myelin sheath with up to 100 layers of membrane only serves as the insulating shell. The speed of action potentials is measured to be as fast as 100 m/s on myelinated axons, but ions move at just 100 nm/s in a 1 V/m electric field. We show here the action potentials, in the form of electromagnetic (EM) pulses, can propagate in natural EM waveguide structures formed by the myelin sheath merged in fluids. The propagation time is mainly cost on the duration for triggering EM pulses at nodes of Ranvier. The result clearly reveals the evolution of axons from the unmyelinated to the myelinated, which has remarkably enhanced the propagation efficiency by increasing the thickness of myelin sheath. Key words: action potential; waveguide; myelinated axon; the node of Ranvier; electric synapse; muscle fiber In nature, a high speed for information transmission among different parts of a live creature is one of the crucial factors that determine its survival chance. The information in a biosystem is transmitted either chemically or physically, or by a hybrid combination of both 1. The velocities for diffusion of molecules in a fluid and for directional movement of ions in an external electric field are both very slow 2. For instance, in a dilute electrolyte, experimental data show that under an electric field intensity of 1 V/m ions can only move at a speed about 100 nm/s. By sharp contrast, an EM wave propagates in vacuum at a speed of 3×108 m/s. The transmission of an EM wave can be regarded as transmission of energy, not transport of mass. With limited scattering effects, an EM wave can travel in vacuum or in a uniform dielectric over a long distance without remarkable loss in energy. This is the reason that Hubble Space Telescope can receive images of galaxies at the edge of the universe, and a high-quality optical fiber can efficiently transfer data across the Pacific. Engineers have also made a variety of transmission lines and waveguides for high-efficiency transmission of EM waves 3,4. It is well known that the action potential is generated by the transient ion flows passing through ion channels. Maxwell equations show that such transient ion flows emit EM waves in the forms of electrical pulses in the way similar to dipole antennas. And, the electrical pulses can propagate along certain transmission path without employing physical movement of charge carriers. Therefore, it is a wise choice for the nature to transmit information through EM waves. However, an EM wave is usually strongly scattered in a complex biosystem made of materials with varied phases, densities, dielectric constants, shapes and sizes. The strong scattering processes are sometimes useful for precise imaging 5-7. The inner part of an axon contains many sub-cellular organelles and cytoskeleton structures such as microtubule, microfilament and neurofilament, thus it has a large resistivity for dc current and strongly scatters an EM wave due to its non-uniformity in dielectric constant, shape, size and density of sub-cellular organelles. As a result, the inner part of an axon is not favorable for transmission of electric signals via either dc currents or EM waves. But fortunately, the phospholipid bilayer has a uniform structure. It is the basic framework of the membrane, and such a membrane covers each and every single cell in a biosystem. In this work, we show that by using the membrane as the main frame, EM waveguide structures can be naturally formed in the inner fluidic environment in biosystems, and such EM waveguide structures are efficient for the propagation of action potentials. Results The EM waveguide structure formed by electrolytes and dielectrics Figure 1a schematically shows the way of transmission for a transverse electromagnetic (TEM) wave, in vacuum or a uniform dielectric. Here the electrical field intensity E, the magnetic induction intensity B and wave vector k are perpendicular to each other. Along the transmission path, the energy density for a TEM wave, often characterized by E2, keeps constant in vacuum and decreases slightly in a waveguide over a long distance. a), Schematic diagram of a TEM wave with wave vector of k Figure 1 travelling in vacuum or a uniform dielectric along z direction. b), A model of a fluid-dielectric-fluid sandwich EM waveguide structure formed from one layer of dielectric media between two layers of electrolyte. c), The calculated REM value versus Cbulk, where the frequency of the EM wave is set at 10 kHz, and the permittivity of the central dielectric layer at 3.0. It is well known that a conductor-dielectric (or vacuum)-conductor sandwich structure can form a parallel plate waveguide. Figure 1b shows the model of an electrolyte-dielectric-electrolyte sandwich EM waveguide structure, which is formed by one layer of dielectric between two layers of electrolyte (see Supplementary 1). The calculated reflection coefficient REM value versus the concentration of bulk electrolyte Cbulk is plotted in Figure 1c, where the frequency of the EM wave is arbitrarily set at 10 kHz, and the permittivity of the central dielectric layer at 3.0. In the range of bulk concentration from 0.01 to 1.0 mol/l, REM increases from 0.58 to 0.95. It suggests that the electrolyte-dielectric-electrolyte sandwich structure works well as an EM waveguide, though not as perfect as a metal waveguide. An EM waveguide can also be formed within a confined fluid channel, as shown in Supplementary 2. The fluid-membrane-fluid sandwich structure as an EM waveguide EM waveguide structures can be formed naturally in biosystems with soft materials only. Figure 2a shows a typical example, where a membrane, shown with phospholipid bilayer, is sandwiched between two layers of fluid full of ions at both sides. As shown in Figure 1c, the reflection coefficient REM is calculated to be 0.87 at the concentration of 0.15 mol/l, a typically ion concentration in a live biosystem. Therefore the fluid-membrane-fluid waveguide structure shown in Figure 2a can be considered as a parallel plate waveguide. Ideally a parallel plate waveguide does not have dispersion on frequency for propagation of TEM waves. The characteristic impedance of such a waveguide is a constant dependent only on the geometry and material parameters of the structure. For the TEM mode, as showed in Figure 2b, in the coordinates where the dielectric layer and the fluid layers are perpendicular to x direction and the EM wave is transmitted along z direction, the two fluid layers work as the conductor cladding plates, thus Ey = Ez = 0, By =0, where Ey and Ez are the electrical field intensities along y direction and z direction, respectively; By is the magnetic field intensity along y direction. Figure 2c schematically shows the field lines of E and B in x-y plane when viewing along the z direction. It gives another reason that why the ion channels are embedded in membrane, generating ion current flows perpendicular to the membrane, for the parallel plate waveguide structure prefer to transmit polarized TEM wave with E perpendicular to the membrane. a), A membrane merged in fluids full of ions at both sides is a Figure 2 typical example of electrolyte-dielectric- electrolyte EM waveguide structure. The dashed purple line indicates a propagating EM wave, but the wavelength of the EM wave is not scaled to the thickness of the membrane. Such an EM wave can be generated by a transient ion current passing through an ion channel embedded in the membrane. b), A schematic three dimensional view of a TEM propagating in the fluid-membrane-fluid waveguide structure along z direction, where E is perpendicular to the membrane. c), A schematic diagram of the field lines of E (represented by arrows) and B (represented by dash lines) in x-y plane viewing along z direction. EM coupling between connected membranes Next we discuss the propagation of EM waves in the membrane network. In a biosystem when two membranes are close or connected by proteins, a kind of dielectric softmaterial, as shown in Figure 3a, an EM wave can inject into the “X”-shaped structure from one side of any membrane, and comes out from both membranes at the other side. This is an important concept showing that an EM wave can travel through a network of coupled membrane network, as long as the membrane network is surrounded by electrolytes. Therefore, the propagation of dendritic impulses in a neural system is dictated by this principle. Figure 3 images of coupled membrane waveguide Cross-sectional structures. a), When two membranes are connected in a biosystem, an EM wave can inject into the “X”-shaped structure from one of the two channels at the left side, and comes out through both channels at the right side, when the surrounded with electrolytes. b), Schematic membrane network is cross-sectional structure of a myelinated axon near a node of Ranvier, where the axon is firmly wrapped with a myelin sheath consisted of multilayered membranes. An EM pulse generated at the node of Ranvier can be transmitted in both the membrane of axon and the myelin sheath. c), An electric synapse, where at the central region there is a gap junction, and the two membranes are connected by connexons. The inset shows the membrane. Figure 3b shows the schematic cross-sectional structure of a myelinated axon near a node of Ranvier, where the axon is firmly wrapped with the myelin sheath, multilayered membranes. An axon is conventionally considered as a transmission line and the myelin sheath acts as the insulating sheath. However it is hard to understand that myelin sheath use even more than one hundred layers of membrane for insulation. Dictated as above, an EM pulse generated at the node of Ranvier can be transmitted both in the membrane of axon and the myelin sheath. Therefore, the myelin sheath acts in fact as the dielectric core of the waveguide structure by adding the thickness of the total membrane layers. Figure 3c shows the cross-sectional structure of an electric synapse, where the ends of two opposite axons are coupled with a gap junction. At the gap, with a narrow spacing around 3 nm, the local membranes of two axons are connected by connexons. No ion channels can generate action potentials existing in the gap junction, but experimentally it has been found that action potentials can propagate through the gap junction from one axon to the other without any measurable delay in time. The mechanism of such a transmission cannot be explained clearly by conventional model with ionic current flows. However, following the model presented in this paper, it is a natural result that the action potential is transmitted in the form of the EM pulse which propagates inside the fluid-membrane-fluid waveguide structure of one axon and passes across the connexon-coupled gap junction and continuously propagates inside the fluid-membrane-fluid waveguide of the other axon. The time delay for such an EM-wave propagation across the coupling junction is in the order of 10 fs, therefore is not measurable. Action potentials “saltatory” between nodes of Ranvier on myelinated axons Now we can apply the fluid-membrane-fluid waveguide structure to explain the propagation of action potential between neighboring nodes of Ranvier along a myelinated axon, so-called “saltatory” mechanism, which is not clearly explained under conventional framework with local circuit ionic current flows. Figure 4 shows schematic pictures of such a process in four frames, where the myelinated axon structure is viewed in its cross-sectional cartoon, and the scales of the axon, axon membrane, myelin sheath and spacing between two nodes of Ranvier do not strictly follow the real scales. The center of the axon is colored in light blue, where three nodes of Ranvier are labeled as A, B, and C. The array of narrow arrows schematically represents the electrical field of the EM pulses transmitted through the membrane of the axon and myelin sheath. Figure 4 Schematic cross-sectional cartoon pictures showing the propagation of action potentials along a myelinated axon among three nodes of Ranvier labeled as “A”, “B” and “C”. a), An EM pulse, E is represented by the array of narrow arrows, propagates from left side passing through Node A, via the path of fluid-membrane-fluid waveguide structure, causing excitation of ion channels at Node A, as indicated with a thick red arrow. b), This excitation causes a transient flow of ions thus generates a fresh EM pulse, and triggers the excitation of ion channels at Node B. c) & d), The same process is repeated so that fresh, EM pulses are generated at Node B and Node C, sequentially, and as a result the EM pulse is transmitted from left to right through the membrane of the axon and myelin sheath. In the first frame, Figure 4a, an EM wave propagates from left side passing through Node A, causing excitation of ion channels at this node. The excitation, indicated with a thick red arrow, causes a transient flow of ions perpendicular to the membrane. Thus according to Maxwell Equations, as shown in the second frame Figure 4b, it generates a fresh pulse of EM wave, which is similar to the function of a dipole antenna. The power of this EM pulse is determined by the number and acceleration of ions passing through the channel in the transient process. This fresh EM pulse propagates through the membrane and myelin sheath of the axon, at a speed of ~ 108 m/s, passes through Node B and triggers excitation of ion channels in the membrane at this node, too. Then, as shown in the third frame Figure 4c, a fresh EM wave pulse is generated at Node B, which propagates shortly along the axon and triggers excitation at Node C. Again, the fourth frame Figure 4d shows the repeated EM wave pulses generated at Node C. If the structure of node of Ranvier, the number and distribution of ion channels, and the triggering mechanism at each ion channel are roughly the same 8-10, we can expect that the fresh EM wave pulses generated at different nodes have the same strength and shape. And this has been proven by many experimental reports. The Propagation efficiency for EM waves in different types of axons Next we examine the propagation efficiency for EM waves in myelinated and unmyelinated axons of various dimensions. In a live cell, the skeleton of membrane, the phospholipid bilayer, is usually embedded with certain density of proteins and other molecules, and it consists of intrinsic structural defects. As a result, an EM wave propagating inside the membrane is remarkably scattered, leading to a limited propagation distance. As well as the fluid, full of ions and molecules, is not a conductor as good as the metal. Now we introduce L0 as the efficient propagation length for an EM wave transmitted along an axon waveguide, defined as in z direction. Assume the original described by 𝐸𝐸 = 𝐸𝐸0 𝑒𝑒 −𝑧𝑧 /𝐿𝐿0 . When z = L0, 2L0, and 3L0, E ≈ 37% E0, 14% E0, and 5% electric field intensity E0 attenuates along the propagation path in the way that can be E0, respectively. Therefore, 1-3 L0 is the distance that an EM wave can be transmitted efficiently, beyond which, the energy strength of the propagating EM wave is assumed to be too small to be detected by natural bio-sensors in a biosystem. We have mentioned that the model shown in Figure 1b is similar to that of a parallel plate waveguide. While wrapped on an axon surface, the membrane forms a co-axial geometry, by which the side-leaking effect of EM energy is remarkably eliminated. The unmyelinated axons are usually thicker than myelinated ones. And, experiments show that the thicker the unmyelinated axons, the faster the transmission speed for action potential. Figure 5 The dependence of effective propagation length L0 on the axon structure and diameter. a) & b), Schematic cross-sections of an unmyelinated axon and a myelinated axon, respectively. Their membrane structure, highlighted with a purple frame in each figure, is a phospholipid bilayer with of calculated L0 versus √r of an unmyelinated axon (shown in solid black line) thickness of d0. The inner and outer diameters are defined as r and R. c), Plots and three myelinated axons with Ns of 10 (shown in red line), 20 (shown in blue line) and 40 (shown in green line). d), Plots of calculated L0 versus Ns of two myelinated axons with r = 1 μm (shown in green line) and r = 4 μm (shown in blue line). Figure 5a and Figure 5b show schematically the cross-sections of these two kinds of axons, respectively. Both of their membranes are basically made of phospholipid bilayer with thickness of d0 and two protein layers attached to both sides of the phospholipid bilayer. From reported experimental data, d0 is measured to be around 3 nm, and the thickness of the two protein layers, d1, is estimated to be around 7 nm. In the figures the inner and outer diameters of these axons are defined as r and R, respectively. Hence R ≈ r+d0+d1 for an unmyelinated axon, and R ≈ r+(2Ns+1)·(d0+d1) for myelinated axon, where Ns is the number of sheath layers 11. Based on the waveguide theory, and assuming the protein layer and the phospholipid bilayer have the same dielectric constant, we have calculated the values of L0 versus square root of r for both unmyelinated and myelinated axons. The results speed for action potential linearly increases with √𝑟𝑟, matching excellently with the are plotted in Figure 5c and Figure 5d. It qualitatively shows that the transmission well known experimental data that reveal a square root dependence of “length constant” for neural signal transmission on the diameter of axons 12-15. It also shows, for EM wave transmission a myelinated axon is much more efficient than an unmyelinated axon of the same inner diameter. For instance, the L0 of an unmyelinated axon with r = 400 μm is roughly the same as that of a myelinated axon with 40 layers of sheath, r = 4.0 μm and overall diameter ~ 10 μm. Discussions With a geometric structure similar to that of a metallic parallel plate waveguide, the fluid-membrane-fluid waveguide is in favor of propagating TEM waves. In a metallic parallel plate waveguide, the component of a TEM wave with electrical field parallel to x direction can propagate for a long distance, but its component with electrical field parallel to y direction cannot. However, as EM waves propagate in fluids in a way different from that in a metal, so the characteristic nature of a metallic parallel plate waveguide is not fully applicable in the waveguides shown in Figure 2. In fluids and membranes, EM waves are expected to show frequency dispersion. As shown in Figure 4, the “saltatory” phenomenon observed in myelinated axons is a natural result of propagation of EM pulses between neighboring nodes of Ranvier. It is important to note that, the propagation of an EM pulse along the membrane of axon and the myelin sheath does not require any external bias voltage or directional current of ions along the axon axis either inside or outside the axon. And, this process does not cause the reversal of membrane potential in the internode region. As this region could be as long as 2 mm, indeed the local electric field generated at any node is screened at this distance, therefore the bias for the local circuit ion current proposed in conventional theory does not exist. The results shown in Figure 5c can be considered as a snap shot of such an evolution. In nature the myelinated axons in complex biosystems are developed from unmyelinated axons in simpler biosystems. One layer of membrane for the unmyelinated axon is developed into multilayer of membrane for the myelinated. In the conventional framework of neuroscience, it is hard to explain why the number of membrane layers for the myelin sheath is observed as many as a hundred or even more. However, as shown in Figure 5c and Figure 5d, in the view of EM waveguide, the reason is clear: The thicker the myelin sheath the better the efficiency of the waveguide. As a result, due to this mechanism, the number of myelinated axons a vertebrate can contain in its spine is a thousand times more than that of unmyelinated ones having the same transmission efficiency, and this is much favorable for advanced communications in the whole body. Note that in Figure 5c and Figure 5d, the plotted efficient propagation length L0 is scaled to arbitrary unit (a.u.) due to lack of the exact experimental data. In 1960s, Kanno & Loewenstein conducted experiments on electrical coupling between gland cells. The results showed attenuation of an EM pulse inside a biosystem with a to L0) of 0.8-1.2 mm 16. The well-constructed “length constant” (similar fluid-membrane-fluid waveguides described in this work is a better path for transmission of EM waves than the random gland cell samples Kanno used in the experiment. Therefore, we can expect an L0 be larger than 1 mm in the best myelinated axons. The transmission speed, va, of an action potential signal propagating along an axon is commonly measured to be 1-100 m/s in different kinds of axons. But an EM wave needs only 10 ns to propagate over a 1 m. Therefore almost all the transmission time is allocated to the generation of EM pulses at the ion channels. We define here τ as the time for the rising edge of a single action potential pulse, which measures 0.1-0.5 ms in various experiments. Hence for myelinated axons when va = 10 m/s, it leads to an average spacing Λ of 1-5mm between two neighboring excitation sources. This rough estimation is well consistent with experimental results showing spacing of 0.2-2 mm between neighboring nodes of Ranvier 17. And clearly, as the ion channels on one axon are almost identical so that τ is a constant, in order to increase the exact transmission speed va, the best strategy for an axon is to reduce the number of nodes of Ranvier per unit length, i.e., to increase the average spacing between two neighboring nodes. As a result, it leads to a simple correlation: va ≈ Λ/τ. Finally, in the following we discuss a strong supporting evidence for the proposed model that fluid-membrane-fluid EM waveguide structures are the signal transmission paths in a biosystem. In a live human muscle tissue, when the muscle fibers are ordered to work (e.g. to shrink), an action potential signal is transferred through an axon to the muscle fibers. At the end of an axon, a synapse is directly contacted to the sarcolemma, an excitable membrane of a muscle fiber. Interestingly, the final sub-cellular structures inside the muscle fiber that reacts with the action potential are a number of myofibrils. Each myofibril is firmly surrounded with a sarcoplasmic reticulum (SR), and each SR is connected via a complex network of T tubules of varied lengths with the sarcolemma, as schematically shown in Figure 6a and Figure 6b. The action of each myofibril is triggered by voltage-dependent calcium channels, which are special proteins embedded in the SR membrane and can releases a large number of Ca2+ ions to the myofibril once triggered. According to the conventional theories, the action potential can be applied to the sarcolemma via a synapse. However, if this is the case, the transmission of electrical signals inside a muscle fiber faces difficulties: (1), there is no axon or excitable membrane connecting the myofibrils and the sarcolemma; (2), there is no loop for local current flow or pressure gradient, i.e., there is no driving force for the motion of ions or molecules in a T tubule; (3), the number of synapse connected to the muscle fiber is much less than the number of myofibril inside each muscle fiber, so that the length of T tubules from each myofibril to the sarcolemma is different from each other, and (4), it is known that the diffusion speed of molecules in solution, and the motion of ions under an external dc field, are both very slow. Figure 6 a), Schematic diagram of a muscle fiber, where a synapse is connected to the sarcolemma. The myofibrils inside the muscle fiber are connected to the sarcolemma via a complex network of T tubules, and each (SR). b), A myofibril reticulum is surrounded with a sarcoplasmic cross-sectional diagram of the black circle in a), consisting of synapse, sarcolemma, T tubule, SR and myofibril. c), A cross-sectional diagram of the black circle in b), showing an electrolyte fluid-phospholipid bilayer-fluid waveguide structure. But in a live biosystem a bundle of myofibrils in a single muscle fiber do react together to the same external action potential signal within a time scale of 20-100 ms 18. This strongly indicates that the same single action potential signal is received by each and every myofibril almost simultaneously. So a consisted explanation is that the signal is not transmitted by molecule diffusion or ion transportation in the T tubules, but by EM waves through the membrane of T tubule. The membrane of T tubule serves as a good EM waveguide, and as a result, although the absolute distance of different myofibrils away from the same synapse-sarcolemma junction varies much, the difference in the transmission times of EM wave is much less than 1 ns, therefore this delay is negligible. Conclusion Clearly, we have shown that to transmit information in a biosystem by EM waves through the fluid-membrane-fluid waveguide structures is physically feasible and an energy-efficient approach. For transmitting an EM wave in a uniform membrane, it needs only a proper source of the EM pulses, such as a transient ionic current flow passing through an ion channel; it does not require any external bias voltage or directional current. The energy dissipation during the propagating of an EM wave is much smaller than that in case of a mass transport. References 1. Debanne, D. Information Processing in the Axon. Nature Reviews Neuroscience 5, 304-316 (2004). 2. Vincent, C. A. The Motion of ions in Solution under the Influence of an Electric Field. Journal of Chemical Education 53, 490-493 (1976). Pozar, D. M. Microwave Engineering (John Wiley & Sons, 3rd Ed., MA, USA, 2005). Pfeiffer, F., David, C., Burghammer, M., Riekel, C. & Salditt, T. Two-Dimensional X-ray Waveguides and Point Sources. Science 297, 230-234 (2002). 3. 4. 7. 8. 9. 5. Chapman, D., Thomlinson, W., Johnston, R. E., Washburn, D., Pisano, E., Gmür, N., Zhong, Z., Menk, R., Arfelli, F. & Sayers, D. Diffraction enhanced X-ray imaging. Physics in Medicine and Biology 42, 2015-2025 (1997). 6. Momose, A., Takeda, T., Itai, Y. & Hirano, K. Phase-contrast X-ray computed tomography for observing biological soft tissues. Nature Medicine 2, 473-475 (1996). Bilderback, D. H., Hoffman, S. A. & Thiel, D. J. Nanometer spatial resolution achieved in hard X-ray imaging and Laue diffraction experiments. Science 263, 201-203 (1994). Ellisman, M. H. Molecular Specializations of the Axon Membrane at Nodes of Ranvier are not Dependent Upon Myelination. Journal of Neurocytology 8, 719-735 (1979). Poliak, S. & Peles, E. The Local Differentiation of Myelinated Axons at Nodes of Ranvier. Nature Reviews Neuroscience 4, 968-980 (2003). 10. Lai, H. C. & Jan, L. Y. The Distribution and Targeting of Neuronal Voltage-gated Ion Channels. Nature Reviews Neuroscience 7, 548-562 (2006). 11. Michailov, G. V., Sereda, M. W., Brinkmann, B. G., Fischer, T. M., Haug, B., Birchmeier, C., Role, L., Lai, C., Schwab, M. H., Nave, K. –A., Axonal Neuregulin-1 Regulates Myelin Sheath Thickness. Science 304, 700-703 (2004). 12. A. L. Hodgkin & A. F. Huxley, The Components of Membrane conductance in the Giant axon of Loligo, Journal of Physiology 116 , 473-496, 1952 13. Hodgkin, A. L. & Huxley, A. F. The Dual Effect of Membrane Potential on Sodium Conductance in the Giant Axon of Loligo. Journal of Physiology 116 , 497-506 (1952). 14. Hodgkin, A. L. & Huxley, A. F. A Quantitative Description of Membrane Current and its Application to Conduction and Excitation in Nerve. Bulletin of Mathematical Biology 52, 25-71 (1990). 15. Goldstein, S. S. & Rall, W. Changes of Action Potential Shape and velocity for changing core conductor Geometry. Biophysical Journal 14, 731-757 (1974). 16. Kanno, Y. & Loewenstein, W. R. Low-resistance Coupling between Gland Cells. Some Observations on Intercellular Contact Membranes and Intercellular Space. Nature 201, 194-195 (1964). 17. Bear, M. F., Cannors, B. W. & Paradiso, M. A. Neuroscience: Exploring the Brain (Lippincott Williams & Wilkins Inc., 2nd Ed., Philadelphia, USA, 2001). 18. Burke, R. E., Levine, D. N., Tsairis, P. & Zajac, III., F. E. Physiological types and histochemical profiles in motor units of the cat gastrocnemius. Journal of Physiology 234, 723-748 (1973). Acknowledgements We thank W.Q. Sun, J. Tang, C.K. Ong, Bonnie W. Lu, and R.J. Dai for valuable discussions. This work is financially supported by Peking University, the NSFC (Grant No 11074010) and MOST of China (Grant No 2011CB933002). Affiliations Department of Electronics, and Key Laboratory for the Physics & Chemistry of Nanodevices, School of Electronics Engineering and Computer Science, Peking University, Beijing 100871, People’s Republic of China. Contributions S.Y.X. proposed the concept that an axon works like a fluid EM waveguide and the EM pulse model for the saltatory mechanism on myelinated axons. J.W.X. found the fluid-phospholipid bilayer-fluid waveguide structure and completed the theoretical calculation. J.W.X. and S.Y.X. wrote the manuscript together. Competing financial interests The authors declare no competing financial interests. Corresponding author Correspondence to: Shengyong Xu, E-mail: [email protected] Supplementary 1: The reflection coefficient REM of EM waves reflected at the smooth interface of in a uniform dielectric In a uniform media, the Maxwell equations for an EM wave can be written in the ∇ × 𝑬𝑬 = − 𝜕𝜕𝑩𝑩𝜕𝜕𝜕𝜕 = 𝑖𝑖𝜔𝜔𝜔𝜔𝜔𝜔0𝑯𝑯 (1) form of the following: ∇ × 𝑯𝑯 = 𝜕𝜕𝑫𝑫𝜕𝜕𝜕𝜕 + 𝑱𝑱 = −𝑖𝑖𝜔𝜔𝜔𝜔𝜔𝜔0𝑬𝑬 + 𝑱𝑱 (2) ∇ ∙ 𝑬𝑬 = 𝜌𝜌𝜔𝜔𝜔𝜔0 (3) ∇ ∙ 𝑯𝑯 = 0 (4) where E, B, H, D and J are vectors for the electric field intensity, magnetic induction intensity, magnetic field intensity, electric displacement and conduction current density; ρ, t, and ω are the free charge density, time and circular frequency of the EM wave; ε0, ε, μ0 and μ are the permittivity of vacuum and the media, the permeability of media with ρ = 0 can be written as 𝑬𝑬(𝒛𝒛, 𝜕𝜕) = 𝑬𝑬𝑒𝑒 𝑖𝑖 (𝒌𝒌∙𝒛𝒛−𝜔𝜔𝜕𝜕 ) = 𝑬𝑬𝑒𝑒 −𝜶𝜶∙𝒛𝒛𝑒𝑒 𝑖𝑖 (𝜷𝜷∙𝒛𝒛−𝜔𝜔𝜕𝜕 ) , where k vacuum and the uniform media, respectively. In general, an EM wave propagating in a �𝛽𝛽2 − 𝛼𝛼 2 = 𝜔𝜔2 𝜔𝜔𝜔𝜔0 𝜔𝜔𝜔𝜔0 (5) = iα + β, with α and β being the attenuation constant and phase constant described as: 𝜶𝜶 ∙ 𝜷𝜷 = 12 𝜔𝜔𝜔𝜔𝜔𝜔0𝜎𝜎 (6) For a plane EM wave incident perpendicularly to a conductor, 𝛼𝛼 2 = 12 𝜔𝜔2 𝜔𝜔𝜔𝜔0 𝜔𝜔𝜔𝜔0 ��1 + 𝜎𝜎 2 𝜔𝜔 2 𝜔𝜔02𝜔𝜔2 − 1� (7) 𝛽𝛽2 = 12 𝜔𝜔2 𝜔𝜔𝜔𝜔0 𝜔𝜔𝜔𝜔0 ��1 + 𝜎𝜎 2 𝜔𝜔 2 𝜔𝜔02𝜔𝜔2 + 1� (8) where σ is the conductivity of the conductor. In vacuum, α = 0. 𝒆𝒆𝒏𝒏 × (𝑬𝑬𝑩𝑩 − 𝑬𝑬𝑨𝑨) = 0 (9) For an EM wave incident to a surface from material A to material B, 𝒆𝒆𝒏𝒏 × (𝑯𝑯𝑩𝑩 − 𝑯𝑯𝑨𝑨) = 𝑱𝑱𝒔𝒔𝒔𝒔𝒔𝒔𝒔𝒔𝒔𝒔𝒔𝒔𝒆𝒆 (10) 𝒆𝒆𝒏𝒏 ∙ (𝑫𝑫𝑩𝑩 − 𝑫𝑫𝑨𝑨 ) = 𝛴𝛴𝑠𝑠𝑠𝑠𝑟𝑟𝑠𝑠𝑠𝑠𝑠𝑠𝑒𝑒 (11) 𝒆𝒆𝒏𝒏 ∙ (𝑩𝑩𝑩𝑩 − 𝑩𝑩𝑨𝑨) = 0 (12) where EA, HA, DA and BA, are the electrical field intensity, magnetic field intensity, electric displacement and magnetic induction intensity in material A, respectively, and EB, HB, DB and BB are the parameters for material B. Jsurface is the surface free current density, Σsurface the surface free charge density, and en the normal vector pointing to the inside of material B. Defined E, E’ and E’’ are the electrical field intensities for the incident, reflected and transmitted EM waves, respectively. So EA = E+E’, and EB = E’’. The reflection coefficient REM is defined as: 𝑅𝑅𝐸𝐸𝐸𝐸 = �𝑬𝑬′𝑬𝑬 �2 (13) An ideal waveguide has REM of 1, while for practical waveguides REM < 1.0, i.e., part of the EM wave energy is lost in the conductor by Joule heating, so that the strength of the EM wave decreases along the transmission path. And, the EM wave energy is also dissipated via scattering process in the dielectric. It has been proven that an EM wave can only penetrate a thin layer of metal. The thickness of this thin layer is defined as the penetration depth δ, and for a good conductor 𝛿𝛿 ≈ � 2𝜔𝜔𝜔𝜔 𝜔𝜔 0 𝜎𝜎 . As a result, an incident EM wave completely reflects at the surface of a perfect conductor, similar to an incident light reflected by a mirror. Thus an EM wave can be trapped between two walls of conductors. A conductor-dielectric (or vacuum)-conductor sandwich structure can form a waveguide, so that an EM wave can travel inside the dielectric layer with little energy leakage. The efficiency of a waveguide for confining the energy of an EM wave in its inner space can be characterized by REM. An ideal waveguide has REM of 1, while for practical waveguides REM < 1.0, i.e., part of the EM wave energy is lost in the conductor by Joule heating, so that the strength of the EM wave decreases along the transmission path. And, the EM wave energy is also dissipated via scattering process in the dielectric media. Along the transmission path, the energy density for a plane EM wave, often characterized by E2, keeps constant in vacuum and decreases slightly in a waveguide over a long distance. According to the Nernst-Planck Equation, the conduction current density J of the 𝐽𝐽(𝑧𝑧) = � 𝑁𝑁𝐴𝐴 𝑒𝑒0 𝑍𝑍𝑖𝑖 � −𝐷𝐷𝑖𝑖 𝜕𝜕𝐶𝐶𝑖𝑖 (𝑧𝑧)𝜕𝜕𝑧𝑧 − 𝐹𝐹𝑅𝑅𝑅𝑅 𝐷𝐷𝐶𝐶𝑖𝑖 𝜕𝜕𝜑𝜑𝜕𝜕𝑧𝑧 + 𝐶𝐶𝑖𝑖 𝑉𝑉(𝑧𝑧)� (14) electrolyte fluid layer along z direction can be written as: i where NA is Avogadro’s constant, e0 being the proton charge; Di, Zi, and Ci are parameters for the ith kind of ions, namely the coefficient of diffusion, valence and concentration, V(z) is the fluid velocity along z direction; F being Faraday constant, R the gas constant, φ the electrical potential, and T the absolute temperature. We consider an electrolyte solution in which the concentrations of cations and anions are the same, defined as Cbulk. If there is no gradient of concentration in z direction and 𝜎𝜎𝑏𝑏𝑠𝑠𝑏𝑏𝑏𝑏 (0), is estimated to be 𝜎𝜎𝑏𝑏𝑠𝑠𝑏𝑏𝑏𝑏 (0) = 2𝜔𝜔𝐸𝐸 𝑁𝑁𝐴𝐴 𝑒𝑒0𝐶𝐶𝑏𝑏𝑠𝑠𝑏𝑏𝑏𝑏 , where μM is the mobility of V(z) = 0, in a uniform applied electric field E, the dc conductivity of bulk electrolyte, ions 1,2. 𝜎𝜎ac (𝜔𝜔) = 𝜎𝜎1 (𝜔𝜔) + 𝑖𝑖𝜎𝜎2 (𝜔𝜔) (15) The ac conductivity σ(ω) of a bulk electrolyte fluid can be described as 3: 𝜎𝜎(0) 𝜎𝜎1 (𝜔𝜔) = 1 + (𝜔𝜔𝐷𝐷𝐷𝐷/𝑏𝑏𝐵𝐵 𝑅𝑅)2 − [2𝜔𝜔2 − 𝜔𝜔4 /𝛺𝛺2 ]/𝛺𝛺2 (16) 𝜎𝜎2 (𝜔𝜔) = 𝜔𝜔𝜎𝜎(0)[𝐷𝐷𝐷𝐷/𝑏𝑏𝐵𝐵 𝑅𝑅 − 𝑏𝑏𝐵𝐵 𝑅𝑅(𝛺𝛺2 − 𝜔𝜔2 )/𝐷𝐷𝐷𝐷(𝛺𝛺2 )2 ] (17) 1 + (𝜔𝜔𝐷𝐷𝐷𝐷/𝑏𝑏𝐵𝐵 𝑅𝑅)2 − [2𝜔𝜔2 − 𝜔𝜔4 /𝛺𝛺2 ]/𝛺𝛺2 where Ω is the Einstein Frequency, kB being Boltzmann’s constant. Only the conduction current, which has the same phase with that of the electric field, consumes the energy. Supplementary 2: An EM waveguide structure consisting of two heavily charged dielectric layers and a fluid with a large gradient of ion concentration Figure S1 a), Schematic model of another possible fluid EM waveguide, where a fluid layer is trapped between two dielectric layers, and a large gradient of ion concentration exists from the edge of the fluid towards the center. b), A plot of calculated REM value versus surface charge density Σs for the waveguide model proposed in a), where Cbulk is set at 1.0×10-7 mol/l. Briefly, in a practical system, when a solid surface is in contact with electrolyte, it is often charged with a surface charge density of Σs. In a dilute electrolyte solution, the average concentration of counterions in the electrical double layer (EDL) 4 is much larger than that in the bulk solution, so a large gradient of ion concentration exists from the edge of the channel towards the center. The thickness of an EDL can be estimated to be the Debye length, 1/κ, where: 𝜅𝜅 2 = 2𝑒𝑒02𝑁𝑁𝐴𝐴 𝐶𝐶𝑏𝑏𝑠𝑠𝑏𝑏𝑏𝑏 𝜔𝜔0 𝜔𝜔𝑏𝑏𝐵𝐵 𝑅𝑅 (18). As a result, the EDL has a much higher conductivity when compared to the dilute bulk solution, and it can serve as the conducting cladding layer of a waveguide. The conductivity of EDL, σEDL can be revealed from Σs or the zeta-potential, ζ 5. Figure S1b shows that, this kind of fluidic waveguide only works well under the condition when the surface charge density is high and the ion concentration of bulk solution is low. By contrast, our calculation shows that when Cbulk is bigger than 0.1mol/l, the EDL phenomenon has negligible effect on REM for the structure shown in Figure 1b. In this waveguide structure, a large gradient of ion concentration exists from the edge of the channel towards the center. In practical systems, when a solid surface is in contact with electrolyte, it is often charged with a surface charge density of Σs. In a dilute electrolyte solution, the average concentration of counterions in the EDL is much larger than that in the bulk solution, as shown in Figure S1a, where the channel surface facing the electrolyte is assumed negatively charged. The thickness of an EDL can be estimated to be the Debye length, 1/κ. Compared to the dilute bulk solution, the EDL has a much higher conductivity, and it can serve as the conducting cladding layer of a waveguide. Thus the fluidic channel structure shown in Figure 1c is indeed an EM waveguide. The conductivity of EDL, σEDL can be revealed from the surface charge density Σs or the zeta-potential, ζ. When the surface charge density is high and the ion concentration of bulk solution is low, the waveguide works well. Figure S1b shows an extreme example, where Cbulk is set at 1.0×10-7 mol/l and REM is plotted versus Σs. In this case, one sees that REM increases with Σs, from 0.54 at 0.01 e/nm2 to 0.95 at 0.1 e/nm2. References for the Supplementary 1. 3. Levine, S., Marriott, J. R. & Robinson, K. Theory of Electrokinetic Flow in a Narrow Parallel-plate Channel. Journal of the Chemical Society, Faraday Transaction 2: 71, 1-11 (1975). Bocquet, L. & Charlaix, E. Nanofluidics, from bulk to interfaces. Chemical Society Review 39, 1073-1095 (2010). Chandra, A., Wei, D. Q. & Patey, G. N. The frequency dependent conductivity of electrolyte solutions. The Journal of Chemical Physics 99, 2083-2094 (1993). Schoch, R. B., Han, J. Y. & Renaud, P. Transport phenomena in nanofluidics. Review of Modern Physics 80, 839-883 (2008). 5. Kirby, B. J. & Hasselbrink Jr., E. F. Zeta potential of microfluidic substrates: 1. Theory, experimental techniques, and effects on separations. Electrophoresis 25, 187-202 (2004). 2. 4.
1905.02007
1
1905
2019-05-06T12:57:12
Axonal Computations
[ "q-bio.NC" ]
Axons functionally link the somato-dendritic compartment to synaptic terminals. Structurally and functionally diverse, they accomplish a central role in determining the delays and reliability with which neuronal ensembles communicate. By combining their active and passive biophysical properties, they ensure a plethora of physiological computations. In this review, we revisit the biophysics of generation and propagation of electrical signals in the axon, their complex interplay, and their rich dynamics. We further place the computational abilities of axons in the context of intracellular and intercellular coupling. We discuss how, by means of sophisticated biophysical mechanisms, axons expand the repertoire of axonal computation, and thereby, of neural computation.
q-bio.NC
q-bio
Axonal Computations Pepe Alcami1,2,∗, Ahmed El Hady3,4,∗ 1. Neurobiology, Ludwig-Maximilians University, Martinsried, Munich, Germany 2. Behavioral Neurobiology, Max Planck Institute for Ornithology, Seewiesen, Germany 3. Princeton Neuroscience Institute, Princeton University, Princeton , New Jersey , USA 4. Howard Hughes Medical Institute, Princeton University, Princeton , New Jersey , USA *correspondence should be addressed to Pepe Alcami ([email protected])or Ahmed El Hady ([email protected]) Abstract Axons functionally link the somato-dendritic compartment to synaptic terminals. Structurally and func- tionally diverse, they accomplish a central role in determining the delays and reliability with which neuronal ensembles communicate. By combining their active and passive biophysical properties, they ensure a plethora of physiological computations. In this review, we revisit the biophysics of generation and propagation of electrical signals in the axon, their complex interplay, and their rich dynamics. We further place the computational abilities of axons in the context of intracellular and intercellular coupling. We discuss how, by means of sophisticated biophysical mechanisms, axons expand the repertoire of axonal computation, and thereby, of neural computation. 1 Introduction Neurons are compartmentalized into input compartments formed by dendrites and somas, and an output com- partment, the axon. However, in the current era, there is a widespread tendency to consider that the biophysics of single neurons do not matter to understand neuronal dynamics and behavior. Neurons are often treated as point processes with disregard of the complex biophysical machinery that they have evolved. Moreover, neuronal computations are assumed to be mostly performed by dendrites or at synapses (Sudhof and Malenka, 2008; Stuart et al., 2016), and axons are reduced to simple, static and reliable devices. However, a wealth of literature supports that this is not the case: axons form complex structures that ensure a variety of sophisticated functions. Here, we aim to review evidence that axons perform complex computations which depend on a myriad of biophysical details and ensure the generation and propagation of neuronal outputs. We will not discuss biophysics of synaptic release, reviewed elsewhere (Sudhof and Malenka, 2008). More than six decades after seminal discoveries in experimentally accessible invertebrate axons (Hodgkin and Huxley, 1952), axonal research has unraveled previously-unsuspected, rich and dynamical electrical signaling in axons, which consists of a hybrid of analog and digital signaling. Contrary to the giant invertebrate axons studied in the early days (Fig. 1A; (Hodgkin and Huxley, 1952; Furshpan and Potter, 1959)), most invertebrate and vertebrate axons are thin and present complex extended arborizations (e.g., Fig. 1B-D), making it difficult to record from them. Electron and optical microscopy have made it possible to deepen our understanding of fine axonal structures (Rash et al., 2016; D'Este et al., 2017). Moreover, the combination of new structural imaging techniques with electrophysiology has made it possible to study how structural changes at the sub-micrometer scale impact function (Ch´ereau et al., 2017). We will first illustrate general signaling principles in axons, before delving into the generation and propagation of action potentials (APs) and their dynamical regulation. Finally, we will position axons in the context of their interactions with other compartments and with each other, by virtue of axo-axonal coupling (Katz and Schmitt, 1940; Furshpan and Potter, 1959) and finally, indirectly via glial cells. 9 1 0 2 y a M 6 ] . C N o i b - q [ 1 v 7 0 0 2 0 . 5 0 9 1 : v i X r a 2 Review of Axonal Computations 2.1 GENERAL PRINCIPLES: SIGNALING IN AXONS 2.1.1 Of axons and brains The propagation of electrical signals along axons controls the reliability and the timing at which neural networks communicate. The spatial extent of axonal trees introduces delays between the generation of an electrical signal (APs or subthreshold signals) in a neuron and the arrival of this signal to the presynaptic site, where information is relayed to a postsynaptic cell. Thereby, the axonal propagation delay influences the temporal relationship of presynaptic and postsynaptic activity (Izhikevich et al., 2004). The delays in AP propagation encode information (Seidl et al., 2010) and contribute to re-configuring neural circuits through plastic mechanisms such as spike timing dependent plasticity (Bi and Poo, 1998; Izhikevich et al., 1 2004). Each synapse has its own 'critical window', given by the delay between presynaptic and postsynaptic activity, to induce synaptic plasticity. The speed at which signals travel along axons can vary over four orders of magnitude, from tens of centimeters per second in thin unmyelinated axons to hundreds of meters per second in giant myelinated fibers (Schmidt-Hieber et al., 2008; Xu and Terakawa, 1999)); highlighting the ability of biophysical specializations characterizing different axons to conduct APs at different speeds. Axons have evolved different intricate geometries (Fig.1), revealing specific structure-function specializations. Constraints to axonal morphology include (1) spatial constraints: axons occupy large volumes of nervous system, yet they require internal components to their function such as mitochondria, which constrain their minimal functional size; (2) energetic requirements may have evolutionary exerted a selection pressure on axonal properties (Perge et al., 2009; Harris and Attwell, 2012); (3) efficiency of electrical information processing in relation to structure-function specializations. The last point on electrical information processing will be the main focus of the current review. 2.1.2 Propagation of electrical signals in the axon Propagation of electrical signals in the axon results from a combination of specialized active and passive mech- anisms. Active properties are shaped by neuronal voltage-gated ion channels recruited as a function of the dynamics of membrane potential whereas passive properties are determined by the axonal membrane at rest and by axonal geometry. As increasingly appreciated, APs are not just an on/off switch, and electrical signaling in the axon should be regarded as a hybridization of analog and digital signalling (Shu et al., 2006). Furthermore, APs are strongly regulated by the background analog activity provided by subthreshold postsynaptic events. For example, the inactivation of potassium Kv1 channels in the axon initial segment broadens the axonal AP waveform and increases unitary excitatory postsynaptic potentials (EPSP) amplitude in layer V pyramidal neurons (Kole et al. 2007). Note that some neurons do not encode information with APs but only with graded signals (graded potential neurons)(Borst and Haag, 1996) . 2.1.3 Non-electrical signaling Signaling in axons and neurons is usually regarded as purely electrical. However, electrical signals are accom- panied by other biophysical changes in the neuronal membrane. The AP is accompanied by changes in many biophysical properties such as temperature, mechanical membrane properties and optical birefringence (Cohen et al., 1970; Tasaki and Byrne, 1992; Howarth, 1975; Howarth et al., 1975; Abbott et al., 1965; Tasaki and Iwasa, 1982b). For example the changes in the optical properties of the axon during propagation have been at the basis of the label free interferometric imaging of APs in in-vitro systems (Oh et al., 2012; Akkin et al., 2007; Batabyal et al., 2017). Mechanical displacements associated with the AP have been measured in many experimental contexts (Tasaki and Iwasa, 1982a,b; Iwasa and Tasaki, 1980; Hill et al., 1977). These mechanical displacements can be regarded as propagating surface modes that are elicited via the large electrostatic force produced by the AP. In this regard, the AP is an electro-mechanical pulse (El Hady and Machta, 2015). Converging evidence suggests that mechanics play a role in electrical signalling (Tyler, 2012). Moreover, there is mounting evidence that some voltage sensitive channels such as sodium and potassium channels are mechanically modulated locally and that many neurons express mechanically-activated channels (Ranade et al., 2015; Schmidt and MacKinnon, 2008; Schmidt et al., 2012). In addition, mechanical pressure at the soma of high-resistance neurons can induce APs (Alcami et al., 2012), an effect likely to be even stronger at axons, which are highly resistive. Theoretical and biophysical evidence of non-electrical signaling has surprisingly not yet led to a thorough investigation of the functional relevance of these mechanical displacements and other non-electrical signaling modalities to axonal computation. 2.2 BIOPHYSICS OF ACTION POTENTIAL GENERATION 2.2.1 A brief historical perspective on the action potential At the resting, non-excited state, mostly potassium channels are open and the resting potential is, as a conse- quence, close to the reversal potential for potassium, maintained around -70 mV. In 1939, Hodgkin and Huxley published the first trace of an AP recorded from the squid axon using an intracellular electrode where one can see a very clear overshoot. Following this in 1949, the proposal that sodium ions are the main mediator of AP generation was put forward by Hodgkin and Katz (Hodgkin and Katz, 1949). They studied the effect of system- atically varying the concentration of sodium ions and measured its impact on the amplitude of the AP recorded. Subsequently, Keynes managed to show that nerve excitation leads to an increase in the transmembrane flow of sodium ions by tracing the movement of the radioactive isotope N a24 in repeatedly stimulated squid axons (Keynes, 1951). These seminal findings confirmed that sodium ions are the main contributors to AP generation. 2 2.2.2 Variations of action potentials APs are characterized by a width, height and an overshoot magnitude (i.e. referring to the AP height beyond 0 mV). Despite sharing a general biophysical mechanism for initiation, the shape of APs varies in different types of neurons. For example, there are cells that exhibit very narrow APs of a few hundreds of microseconds in width, reflecting a fast spiking behavior, such as some GABAergic interneurons, Purkinje neurons which are GABAergic projection neurons, glutamatergic neurons of the subthalamic nucleus and Medial nucleus of the trapezoid body (MNTB) cells in the superior olivary complex. CA1 pyramidal neurons in the hippocampus have, in contrast, relatively wide AP and dopaminergic neurons have an even wider AP of up to 4 ms ( reviewed in (Bean, 2007)). Apart from the shape of the single spike, neurons can exhibit a diversity of firing patterns. For example, they can be bursting or non-bursting on one hand and they can be adapting or non-adapting on the other hand. Adaptation during a train of APs refers to the process by which AP properties, typically rate and amplitude, decrease within the spike train. There are many biophysical mechanisms by which spike frequency adaptation can happen. The most prominent mechanisms are an increase in outward current flowing through calcium-activated potassium channels and an increasing outward current produced by the electrogenic sodium-potassium pump (Powers et al., 1999). Moreover, there is also a substantial sodium channel inactivation induced by a long lasting depolarization (Sawczuk et al., 1997). The diversity of AP shapes along with the firing patterns are a result of the different combinations of ion channels that the neuron expresses (reviewed in (Marder and Goaillard, 2006)). 2.2.3 The axon initial segment Before delving into the mechanistic details of AP generation, it is crucial to appreciate the complex anatomy of the Axon Initial Segment (AIS), the site of AP initiation (Fig. 3A). Generally speaking, the AIS refers to the first 20-60 µm of the axon but the specific length of the AIS is variable across neurons as it has to adapt it to its own excitability properties. The excitability of the AIS is modulated by varying compositions and subtypes of sodium and potassium channels clustered at the AIS. As an example, motor alpha and gamma neurons have an AIS length of 28 - 29 µm but alpha and gamma motor neurons differ in the percentage of N av 1.6 and N av 1.1 channels subtypes distribution (Duflocq et al., 2011) which underlies the different excitability properties of both subtypes of motor neurons. It is important to note that the different excitable properties are owed to the fact that N av 1.6 and N av 1.1 channels have different kinetics. The AIS contains a highly-specialized protein machinery that gives it a distinct character. One such protein is Ankyrin G, a scaffolding protein that is also present at the nodes of Ranvier (Kordeli et al., 1995). Ankyrin G anchors voltage-gated channels such as voltage-gated sodium (N av) and potassium channels (Kv) to the membrane along with other adhesion molecules (Davis et al., 1996). Beta IV spectrin is another protein expressed in the axon initial segment. Beta IV spectrin's main function is to cluster sodium channels at the axon initial segment while simultaneously binding to the actin cytoskeleton (reviewed in (Rasband, 2010))( Fig.3 A). The AIS in mammalian neurons has been established as the site of the AP initiation following a series of seminal studies that began in the mid 1950s (Araki and Otani, 1955; Coombs et al., 1957; Fatt, 1957). The location of AP initiation has been further confirmed by combining precise electrophysiological measurements and imaging technologies. These allow to precisely identify the locus of AP initiation which was established to be in the distal part of the AIS, 20 to 40 µm from the soma (Atherton et al., 2008; Foust et al., 2010; Kole et al., 2007; Meeks and Mennerick, 2007; Palmer et al., 2010; Palmer and Stuart, 2006). The AIS is considered to be the site of AP initiation also because of several key properties. It contains a high sodium channel density. Furthermore, AIS sodium channels show a voltage dependence shifted to lower voltages, favoring their activation at less depolarized voltages than at the soma. Finally, the relatively large electrotonic distance of the AIS from the capacitive load of the soma renders distal sodium influx more efficient in evoking a local membrane depolarization, compared to an AIS that would start at the soma. Note that electrophysiological recordings from invertebrates have shown that the initiation of the AP can happen at multiple locations acting in an independent manner (Calabrese and Kennedy, 1974; Maratou and Theophilidis, 2000; Meyrand et al., 1992). Along with its specialized protein machinery and acting as the site of AP initiation, the AIS also acts as a diffusion barrier between the somatodendritic and axonal compartments, filtering transport materials passing from soma to axon (Song et al., 2009; Brachet et al., 2010). Note however that the AIS length can be regulated in an activity- dependent manner (Kuba et al., 2014) and that in some neurons, AP initiation has been reported to occur at the first node of Ranvier (Lehnert et al., 2014; Clark et al., 2005) , as will be developped later. 2.2.4 The sodium ionic dynamics and action potential initiation The extent to which the density of sodium channels is higher in the AIS and the contribution of this high density of sodium channels to AP initiation is a matter of active investigation. There is a consensus that sodium channel density is higher in the AIS but the order of magnitude is still unclear. Immunostaining of sodium channels consistently indicates that there is higher channel density in the AIS of various neuron types (Meeks and Mennerick, 2007; Wollner and Catterall, 1986; Boiko et al., 2003). Lorincz and Nusser in (Lorincz and Nusser, 3 2010) counted around 200 N av 1.6 sodium channels per micrometers square in the AIS of hippocampal pyramidal cells using electron microscopy. Given that the conductance of a single N av channel is around 15 pS ( Colbert and Johnston, 1996), one would expect a conductance density of about 3000 pS per micrometers square. On the contrary, electrophysiological measurements from membrane patches pinpoint that sodium channel density is about 3 4 channels per micrometers square in the AIS, which is the same as the somatic density (Colbert and Johnston 1996, Colbert and Pan 2002). A similar conclusion of equal densities was reached on the basis of recordings from blebs that form when cortical axons are cut. This might be due to the inability to draw AIS N a+ channels into the patch-clamp recording pipette due to their tight coupling to the actin cytoskeleton (Kole et al., 2008). In this article, the authors record a much larger sodium current after disruption of the actin cytoskeleton (Kole et al., 2008). Kinetics of sodium currents underlying AP generation have been extensively studied. Hodgkin and Huxley have proposed that the activity of sodium channels can be fitted by m3 activation kinetics. Baranauskas and Martina (2006) (Baranauskas and Martina, 2006) found that sodium currents in three types of central neurons (prefrontal cortical cells, dentate gyrus granule cells and CA1 pyramidal cells) activate faster than predicted by Hodgkin-Huxley type kinetics following m2 activation kinetics. Moreover, it was found that the half activation voltage of voltage-gated sodium channels in layer 5 pyramidal neurons is 7 -14 mV lower in the distal AIS compared to the soma and decreases further with increasing distance from the soma (Colbert and Pan, 2002; Hu et al., 2009). N av1.6 channels, which are predominantly expressed in the axon initial segment and have a lower half activation voltage (Rush et al., 2005), are proposed to be primarily responsible for the initial slope of the AP. Sodium channels at the AIS are more capable of producing a persistent sodium current (Astman et al., 2006; Stuart and Sakmann, 1995). The persistent sodium current has a significant influence on the AP threshold (Kole et al., 2008). Moreover, it is implicated in the generation of the AP afterdepolarization and it therefore contributes directly to the generation of high frequency AP bursts (Azouz et al., 1996). Using sodium imaging to follow sodium influx in axon, soma and basal dendrites (Fleidervish et al., 2010), the authors suggest that the ratios of N av channel densities in these regions are approximately 3:1:0.3. Interestingly in another study by (Lazarov et al., 2018), APs were initiated in the AIS, even when axonal N av channel density was reduced to about 10 percent in a beta IV spectrin mutant mouse. This experimental finding indicates to a great extent that AP initiation in the AIS does not require such a high local channel density. However in that study, the precision of AP timing was substantially compromised when axonal channel density was reduced. Likewise, Kopp-Scheinpflug and Tempel found a decrease in the temporal accuracy of AP generation from MNTB cells in a beta IV spectrin mutant mouse (Kopp-Scheinpflug and Tempel, 2015) The aforementioned section highlights the complexity of sodium ion dynamics and begs for novel imaging modalities that allow tying ultrasfast dynamics with the axonal ultrastructures in order to get insights into the intricate biophysical mechanisms underlying the very first microseconds of AP initiation. 2.2.5 The action potential rapidness A simple but very informative way to study the properties of the APs is to plot the time derivative of the voltage (dV/dt) versus the voltage. This is called phase-plane plot. The spike threshold can be easily visualized in such a representation, where it corresponds to the voltage at which dV/dt rises abruptly. Coombs et al. (Coombs et al., 1957) noticed that the main spike is preceded by a smaller earlier component (referred to as a kink). This component is interpreted as reflecting initiation of the spike in the initial segment of the axon. One can record such a component in somatic spikes in many central neurons, including neocortical pyramidal neurons which we will focus our discussion on here. As mentioned above, one of the most striking features of the AP recorded from cortical neurons has been the existence of a kink at the initiation of the AP (Fig. 3B1). One can define the rapidness of AP onset in this cell as the slope of the phase-plane plot at dV/dt = 10 mV /ms. In (Naundorf et al., 2006), the AP rapidness was measured from the cat visual cortical cells. AP rapidness varied between around 20 to 60 ms−1. It is important to mention that sharp, step-like onsets of APs have been recorded in vivo in many preparations (cat visual cortex: (Azouz and Gray, 1999), and cat somatosensory cortex: in (Yamamoto et al., 1990)). Several hypothesis have been proposed to explain the origin of the AP kink: 1) the backpropagation to the soma of a smoother AP generated at the AIS; 2) an abrupt opening of sodium channels due to the biophysics of neuronal compartmentalization; 3) a decreased membrane time constant due to the loading of the dendritic compartment and 4) the cooperativity of sodium channels at the AIS. The lateral current hypothesis states that the kink at spike onset reflects lateral current coming from the axon which becomes sharper through backpropagation from the initiation site to the soma while initiation is smooth at the initiation site (Yu et al., 2008; McCormick et al., 2007) . In Yu et al. 2008, authors perform simultaneous recordings from axon blebs and soma, finding a smoother AP onset in the axon, additionally reproducing these results in a model. It is important to note that this study supporting the lateral current hypothesis was done in blebs recordings which are injured axons that may have undergone severe cytoskeletal reorganization (Spira et al., 2003) affecting sodium channels dynamics. This reorganization might alter the true dynamics of AP initiation. 4 Although the kink indeed might reflect the lateral current coming from the axon (Milescu et al., 2010), this hypothesis fails to account for the ability of neurons to follow 200 - 300 Hz frequency inputs. In order to account for this discrepancy, Romain brette in (Brette, 2013) proposed that the compartmentalization and the distance between the soma and the AIS leads to spike initiation sharpness. In his proposal, Brette suggests that the rapidness arises from the geometrical discontinuity between the soma and the AIS, rather than from the backpropagation of axonal APs. When sodium channels are placed in a thin axon, they open abruptly rather than gradually as a function of somatic voltage, as an all-or-none phenomenon. Another proposal that takes into account the geometry of neurons is Eyal et al. (Eyal et al., 2014) in which the authors propose that increasing the dendritic membrane surface area (the dendritic impedance load) both enhances the AP onset in the axon and also shifts the cutoff frequency of the modulated membrane potential to higher frequencies. This dendritic size effect is the consequence of the decrease in the effective time constants of the neuron with increasing dendritic impedance load. The authors have shown this in a computational model of reconstructed layer 2 / 3 pyramidal neurons of humans and rats. The firing pattern at the axon is strongly shaped by the size of the dendritic tree. Authors predict that neurons with larger dendritic trees have a faster AP onset. Another proposal to interpret AP onset rapidness is that sodium channels, which are assumed to be opening independently within the Hodgkin-Huxley framework, are gated cooperatively. The cooperativity model proposed that the half-activation voltage of the channels becomes dependent on the probability of the opening of the neighboring channels. These cooperative effects might happen mechanistically on very fast timescales either through a purely electrical, mechanical or electro-mechanical coupling. Though there is no direct experimental test of the cooperativity of neuronal sodium channels at the AIS, it can theoretically account for the observed discrepancy between the sodium channel density and the very rapid rise of the AP at the site of initiation in the initial segment (Naundorf et al., 2006). It is important to note that cooperative gating has been previously observed in calcium, potassium and HCN channels (Marx et al., 2001; Kim et al., 2014; Dekker and Yellen, 2006). Apart from its mechanistic underpinnings, the fast rise of APs has attracted both experimental and theoretical approaches to study its functional implications on the biophysics of neuronal populations. Before detailing those functional implications, it is important here to go through a useful theoretical abstraction: a typical cortical neuron, embedded in a cortical network in vivo, receives about 10.000 synaptic inputs. Assuming that each of these synaptic inputs is active with a rate on the order of 1 to 10 Hz, incoming signals arrive at a rate of 10 kHz. As a result, the membrane voltage exhibits strong, temporally irregular fluctuations. To understand the computational capabilities of e.g. cortical circuits, it is essential to characterize single neuron computation under such realistic operating conditions. To control the activity of entire neuronal circuits while preserving their natural firing characteristics, it would be advantageous to introduce artificial input components mimicking intrinsically generated synaptic input under precise experimental control. In order to study the dynamical properties of cortical neurons, experimenters have mimicked synaptic bombardment in vitro by injecting stochastic inputs modeled as an Ornstein-Uhlenbeck process in which sinusoidal inputs are embedded (Neef et al., 2013; Tchumatchenko et al., 2010). This experimental setting has allowed the measurement of the dynamic gain of neurons, which means how much neurons attenuate their input in the frequency domain and how fast they are able to follow a rapidly-fluctuating input. This has led to the establishment of the ability of cortical neurons to follow high frequency inputs up to 200-300 Hz (Higgs and Spain, 2009; Higgs et al., 2006; Tchumatchenko et al., 2011). Moreover, there has been a series of theoretical studies exploring the dependence of encoding capacity on the active properties of the AP initiation (Fourcaud-Trocm´e et al., 2003; Huang et al., 2012; Wei and Wolf, 2011) . 2.2.6 AP trajectories beyond the rapidness of initiation Although we have mostly concentrated initial spike rapidness, note that the phase plot can display a variety of trajectories after the onset of the AP (Fig.3). These trajectories are determined by additional ion channels in conjunction with sodium channels. Potassium channels are typically responsible for the repolarizing phase of the AP. The relative temporal profiles of activation of sodium and potassium channels and their subunit composition determine the width of the AP (Lien and Jonas, 2003). Furthermore, the spatial location of ion channels also contributes to the shape of AP trajectories (Yang et al., 2016)(Fig. 3B2). It has also been shown that the AP shape affects calcium currents and transmitter release in a calyx-type synapse in the rat auditory brainstem (Borst and Sakmann, 1999). Therefore, the AP shape beyond the initial spike rapidness provides an extra dimension for information encoding on a variety of timescales. 2.2.7 The action potential at nodes of Ranvier Although the AP is generated at the AIS, the nodes of Ranvier contain a machinery to regenerate the AP. It is important to note that there are striking similarities between axon initial segment and nodes of Ranvier. A great deal of the protein machinery in the axon initial segment is also present in the nodes of Ranvier where the regeneration of the AP is performed. Ankyrin G, the major scaffolding protein in the AIS and the ring-like arrangement of actin and beta IV spectrin are also found in the nodes of Ranvier (DEste et al., 2017). Nodes 5 of Ranvier are distributed spatially along the axon to guarantee the faithful propagation of the AP. In addition, the first node of Ranvier was found to be crucial for high bandwidth bursting activity in neocortical layer 5 pyramidal neurons (Kole, 2011). In that study, authors show that nodal persistent sodium currents at the first node of Ranvier hyperpolarize AP threshold and amplify the afterdepolarization. The study opens up the space for a computational role of the first node of Ranvier beyond conduction of the propagating AP. The first node faithfully follows spike frequencies with a 100 µs delay (Foust et al., 2010; Khaliq and Raman, 2006; Palmer et al., 2010; Palmer and Stuart, 2006). This had led to the speculation that the nodes of Ranvier, and in particular the first node, may have an active computational role in modulating the AP initiation itself. It is worth noting that nodes of Ranvier also express N aV 1.6, the same sodium channel subtype that is ex- pressed in the AIS. It is even more striking that the sodium channels at the nodes also undergo the developmental changes from N aV 1.2 to N aV 1.6 during postnatal development, following a similar developmental trajectory as those found in the AIS (Rios et al., 2003). N aV is highly clustered at the nodes of Ranvier in the order of 1200 channels per micrometers square (Rosenbluth, 1976), while internodes contain 2025 channels per micrometers square (Ritchie and Rogart, 1977)). This very high density ensures high-fidelity conduction of the AP which will be further developed in the review. Given the close proximity of the first node to the cell body and high density of N aV channels, it has been postulated that, in addition to securing propagation, it could potentially generate the AP (Clark et al., 2005; Colbert and Pan, 2002; Lehnert et al., 2014) Although this might be happening, the overwhelming evidence favors that the AP initiation is happening at the AIS. 2.2.8 Ectopic spiking AP initiation at the AIS and its subsequent orthodromic propagation have been extensively investigated. How- ever, a number of studies demonstrates that distally generated spikes or ectopic APs co-exist in invertebrate and vertebrate neurons (Mulloney and Selverston, 1972; Maranto and Calabrese, 1984; Meyrand et al., 1992; Sheffield et al., 2010; Dugladze et al., 2012; Lehnert et al., 2014). Marked differences in somatic AP recordings are observed when those are generated at the AIS or in a distal part of the axon. In particular, APs have more negative thresholds when they are generated in distal parts of the axon. This is likely due to the fact that they do not show the strong coupling of soma to AIS which is responsible for the inactivation of sodium channels at the AIS and for a more depolarized AP generation threshold. Interestingly, given the different potential at which ectopic spikes are generated, the conductances activated during AP generation may differ between distal and proximal APs (Meyrand et al., 1992). In invertebrates, numerous examples of ectopic spikes and of their functional relevance have been described (Mulloney and Selverston, 1972; Maranto and Calabrese, 1984; Meyrand et al., 1992). Ectopic spikes in neurons that target the hearts of the leech have been proposed to control the heart firing frequency (Maranto and Calabrese, 1984). In the somatogastric ganglion of the crab, ectopic spikes are typically observed in the lateral gastric motor neuron only when the muscles remain attached to the preparation, ectopic spikes being induced by motor contraction (Meyrand et al., 1992). Remarkably, ectopic spikes in (Meyrand et al., 1992) fail to depolarize terminals onto interneurons located close to the soma, whereas they efficiently excite a distal postsynaptic target, the muscle. Thus, orthodromically-propagating spikes generated close to the soma and antidromically- propagating spikes generated distally co-exist and they can reach synapses that target different postsynaptic neurons. Ectopic spikes have also been observed in vertebrates. In principal cells from CA3 area in the hippocampus during rhythmic activity in the gamma range (Dugladze et al., 2012), axons fire APs at five times the firing frequency detected at the soma. This is due to the activity of a specific interneuron, the axo-axonic cell, that inhibits the initial segment, thereby avoiding the backpropagation of axonal spikes to the soma. Another noteworthy case of co-ocurrence of AIS and ectopic spikes has been reported in (Lehnert et al., 2014) in auditory medial superior olive (MSO) neurons. A realistic model of MSO neurons which takes into account their axonal structure and ion channel composition together with a physiological synaptic bombardment regime generated APs at both the AIS and at the first node of Ranvier. Additionally, under certain pathological conditions, e.g. in epilepsy, hyperexcitable axons have been reported to generate ectopic spikes in the hippocampus (Stasheff et al., 1993). We would like to make at this point a clarification: although 'action potential' and 'spike' are two terms used to refer to the sodium AP generated at the axon initial segment, the term spike seems preferentially used in the literature to refer to spikes that can differ in their location (dendritic or ectopic spikes) and in their underlying ionic mechanism (e.g., calcium spike, see below). 2.2.9 Beyond the sodium spike The spike initiated through sodium influx is not the only spike propagating in the axon. There are spikes that are generated through the influx of calcium. In the giant axon of the jellyfish Aglantha digitale (order Hydromedusae), both sodium and calcium spikes propagate in the axon (Mackie and Meech, 1985). Sodium- 6 dependent spikes are responsible for fast swimming and calcium spikes mediate slow swimming. Also in the axon giant neurone R2 of Aplysia, the propagating AP has a mixed Na/Ca dependency (Horn and Miller, 1978) . In vertebrates, calcium spikes are typically restricted to the dendritic compartment where there is a calcium spike generation mechanism. The biophysical mechanisms contributing to the dendritic spikes initiation in the distal apical trunk and proximal tuft of hippocampal CA1 pyramidal neurons (Gasparini et al., 2004) are (1) dendritic sodium and potassium channels that set the threshold and determine the shape and forward-propagation of dendritic spikes, (2) the highly synchronized inputs (approximately 50 synapses activated within 3 msec) and (3) the spatial clustering (all inputs arriving in 100 µm of the dendrite). Moreover, NMDA receptors enhance spike initiation by counteracting the shunting of AMPA synaptic conductances. In the olfactory bulb mitral cells and in hippocampal and cortical pyramidal cells, dendritic spikes can trigger one or more axonal APs (Chen et al., 2002; Stuart et al., 1997; Golding and Spruston, 1998; Larkum et al., 1999, 2001; Ariav et al., 2003) . In addition, backpropagating axonal APs can themselves promote dendritic spikes, a reciprocal interaction that can lead to a burst of axonal APs (Pinsky and Rinzel, 1994; Mainen and Sejnowski, 1996; Larkum et al., 1999, 2001; Doiron et al., 2002). 2.3 BIOPHYSICS OF ACTION POTENTIAL PROPAGATION 2.3.1 An equivalent electrical circuit for axons The passive properties of axons can be modeled by an equivalent electrical circuit (Fig. 3A, top). The axonal membrane can be reduced to two circuit elements: the lipid bilayer, modeled by a capacitor and ion channels, by a resistor. The resistance to axial current flow can be modeled by an additional resistance. How fast signals propagate is critically controlled by the capacitance. Electrical currents need to first charge the membrane capacitance, that opposes the flow of electric current, before electrically-charged membranes can undergo voltage changes. Capacitances are proportional to the membrane capacity (capacitance per surface area) and to the membrane surface area. The capacity of neuronal membranes is in the order of 1 µF per cm2 (Gentet et al., 2000). Capacitance measurements from small axons are in the range of tens of picoFahrads (pF) (Mejia-Gervacio et al., 2007). Capacitances of larger neurons (i.e., strongly-ramifying interneurons and projection neurons) are therefore expected to be in the range of hundreds of pF to the nF range. Dynamic changes in capacitance are suggested by activity-dependent changes in the size of axons (e.g.,(Ch´ereau et al., 2017)). Little is known about dynamical changes in capacitance due to temperature, lipid composition, and whether these may significantly impact the propagation of electrical signals along the axon. The membrane resistance plays a major role in controlling how far signals spread in space along the axon before membrane potential changes become imperceptible. This is typically measured by the space constant, defined as the distance over which the membrane potential decays to 37 percent of its initial value. The more open channels are available at the membrane, the more the axial current is attenuated in space along the axon due to current leakage through membrane channels. Remarkably, the membrane resistance can change (reviewed in (Debanne et al., 2019)), providing a potential source of variation in the conduction of electrical signals along axons. The axonal axial resistance to current flow results from the combination of the resistivity of the axoplasm (that is, the cytoplasm of the axon), given in Ohms*cm, and the diameter of the cylinder. Axial resistivity values are in the range of 100 Ω.cm (Carpenter et al., 1975; Cole, 1975). Computational models of the calyx of Held suggest that the latency and amplitude of signals propagating between release sites is highly sensitive to changes in axial resistivity, and that these changes may have a large impact on synaptic release (Spirou et al., 2008). However, changes in resistivity have not been reported so far experimentally. The axon diameter can vary in three orders of magnitude, from around hundred nanometers to hundreds of micrometers in diameter. Giant fibers found in invertebrates are an example of specialized large-diameter structures that propagate electrical signals with high speed over long distances ( in (Hodgkin and Huxley, 1952; Xu and Terakawa, 1999), Fig. 4A). The larger the diameter of the axon, the faster the electrical signal propagates. Taking into account cable theoretic considerations, the propagation speed increases with the square root of the diameter. Remarkably, the axon diameter, as well as the diameter of synaptic boutons, are not static properties of axons. In fact, axon diameter has been shown to be regulated in hippocampal principal cells in an activity-dependent manner. Plasticity protocols induced changes in axonal diameter which were accompanied by significant changes in AP conduction velocity along CA3 pyramidal cell axons (Ch´ereau et al., 2017)(Fig. 5A). 2.3.2 Propagation of subthreshold signals Subthreshold signals propagate passively in axons, reaching synaptic terminals, where they influence synaptic release (Alle and Geiger, 2006). Subthreshold membrane fluctuations consist of synaptic events, typically in the range of a 100s of µVs to mVs, which in cortical neurons approximate highly stochastic background dynamics in vivo (Rudolph and Destexhe, 2003). Their propagation, referred to as analog, in comparison to the digital nature of the AP, modulates the efficiency of APs to induce neurotransmitter release in synaptic terminals (Shu et al., 7 2006). Given that subthreshold signals are not regenerated along the axon, analog signaling is more prominent in proximal portions of the axon. The presence of combined analog and digital signals in axons has been termed hybrid analog-digital signaling. Of particular interest is that the propagation of slow analog subthreshold signals does not only reach chemical synapses. Subthreshold signals also reach axonal electrical synapses, at which the signal is expected to be conveyed with high efficiency to the postsynaptic site due to the continuous and low-pass filtering properties of electrical transmission (reviewed in (Alcami and Pereda, 2019)). 2.3.3 Reliability of propagation and action potential failures Before we consider how signals propagate along axons, let us discuss the reliability of propagation. APs do not always efficiently propagate, and in fact, failures in AP propagation are observed in all neuron types at high firing frequencies (Krnjevic and Miledi, 1959; Grossman et al., 1979; Monsivais et al., 2005). Although geometry can generate failures (e.g. at branch points, see next section), failures typically involve active mechanisms during repetitive activation of axons as observed in a number of vertebrate and invertebrate preparations (Krnjevic and Miledi, 1959; Monsivais et al., 2005; Mar and Drapeau, 1996). Repetitive activity typically leads to extracellular potassium accumulation and subsequently to the depolarization of the axon, inducing failures of conduction (Grossman et al., 1979). In other cases however, repetitive activation induces AP propagation failures via hyperpolarization of the membrane (Mar and Drapeau, 1996). Note that propagation failures likely affect only a fraction of APs at physiological firing rates, and that most APs succeed in propagating. in response to synaptic inputs. Indeed, specific potassium channels of the type IA underlie a hyperpolarization-mediated conduction block in hippocampal pyramidal cells (Debanne et al., 1997). Remarkably, the activation and de- inactivation kinetics of IA allow for a history-dependent conduction block of APs. Interestingly, changes in AP failure can also occur in response to membrane fluctuations, e.g. 2.3.4 More than linear cables: Impact of axonal branching and inhomogeneities Axons are typically formed by complex trees and spatial heterogeneities. These include branching points, vari- cosities (local enlargements of the axon containing the release machinery of chemical presynaptic sites) and large structures specialized in the interaction with other cells. Examples of such structures are the basket formed by a specific type of interneuron, basket cells, around principal cells in many brain regions; the glomerular collateral of the climbing fiber in the cerebellum; the pinceau structure surrounding cerebellar Purkinje cells (Palay and Chan-Palay, 2012) ; Fig. 1B, C) or the calyx of Held in the auditory system. Changes in axon diameter at varicosities or branching points are characterized by an impedance mismatch, that is, a need of a larger current to flow in one of the two directions to electrically load axonal branches, provoking changes in AP propagation (Goldstein and Rall, 1974; Manor et al., 1991). Axonal branches indeed represent a local challenge to the propagation of electrical signals, due to the impedance mismatch between the mother and daughter branches. That is, if APs need to load a larger impedance as they propagate from one branch into two branches, their propagation will be delayed relative to the speed that they would have had if no branching was present. In the most extreme case, the electrical signal fails to load one or two of the branches, resulting in AP propagation failure. Propagation failures have been shown to occur in branches of a number of neurons (Yau, 1976; Gu et al., 1991; Grossman et al., 1979). An interesting example is provided by the medial pressure sensory neuron in the leech, where the failure of APs to propagate can differentially affect postsynaptic cells contacted by distinct presynaptic branches (Gu et al., 1991). Finally, branching points can also provoke a surprising effect: APs can be slowed down in the ms range, up to a level that allows the mother branch to overcome the refractory period for AP generation. As a consequence, the AP can 'reflect' or travel backwards, increasing synaptic release at synapses present in the branch where the AP reflects (Baccus, 1998; Baccus et al., 2000). The impact of complex morphologies on AP propagation is likely to be relevant in the complex axonal ramification patterns of many neurons, including vertebrate interneurons (Ofer et al., 2019). The usage of voltage sensitive-dyes that track membrane voltage with sub-millisecond precision (Palmer and Stuart, 2006) should allow following large portions of axons both in vitro and in vivo, and characterize the propagation of APs along complex axonal structures. 2.3.5 Biophysical properties of myelinated fibers Many vertebrate and some invertebrate fibers are myelinated, a specialization endowed by the glial ensheathment of axons (Castelfranco and Hartline, 2015). Myelin, formed by compact lipidic layers produced by the membranes of glial cells (oligodendrocytes in the central nervous system and Schwann cells in the peripheral nervous system), strongly impacts the propagation of electrical signals. 8 In his seminal study, Lillie (Lillie, 1925) wrapped an iron wire placed in an acidic solution with an insulating glass cylinder. This increased the speed of propagation of electrical signals along the wire, which he postulated to occur in myelinated fibers. His prediction was confirmed decades later in axons by recording electrical signals at nodes, leading to the concept of the saltatory nature of transmission of electrical signals in myelinated fibers (Huxley and Stampfli, 1949). Saltatory comes from the latin verb 'saltare' (to jump), an analogy describing the very fast propagation of electrical signals between nodes of Ranvier. Myelin modifies the electrical circuit that models the passive properties of axons described above. It increases the effective radial resistance and decreases the effective capacitance of the axon (Fig. 4). This effect is due to the different properties of series resistors and series capacitors added to the circuit: series resistors sum their resistances and the inverse of capacitances from capacitors in series sums. The increase in effective membrane resistance (Bakiri et al., 2011) and the decrease in effective membrane capacitance by myelin have two major consequences. On the one hand, the increase in the effective axonal resistance by myelin increases their length constant. On the other hand, as a consequence of the decrease in the effective axonal capacitance, the time required to effectively load axons decreases, dramatically accelerating the propagation of electrical signals. This speeding of electrical propagation by a reduced effective capacitance underlies saltatory conduction between nodes of Ranvier (Huxley and Stampfli, 1949; Castelfranco and Hartline, 2015) . 2.3.6 Geometry of axons and myelin Specific myelination parameters have been shown to maximize conduction velocity of electrical signals along axons. In particular, the ratio of the axonal diameter d to the fiber diameter D (the summed diameter of axon and myelin sheath), defined as the 'g-ratio', has been shown to critically control the conduction of electrical signals. Rushton (Rushton, 1951) developed a biophysical formalism to model current flow in a myelinated axon. He deduced the relation between the ratio l/D (internode length l over D) and the g-ratio. He further demonstrated that the ratio l/D is maximal when g = 0.6, a value also found analytically to maximize the space constant. In an independent approach, Deutsch (Deutsch, 1969) mathematically derived the geometry of axonal and myelin properties that maximize, this time, conduction velocity. He deduced that the propagation velocity is inversely proportional to the internode length and to the RC time constant given by the internal resistance of the axon and the capacitance of the membrane. Maximizing conduction speed requires minimizing the time constant of the circuit. This resulted in the same geometrical properties as those derived by Rushton: a g-ratio of 0.6. Additional modeling studies including a more complete description of myelinated fibers converged to similar conclusions (Goldman and Albus, 1968). Therefore, due to the biophysical constraints posed by the thickness of myelin and axonal size, both conduction speed and efficient spatial propagation of electrical signals are maximized by specific axonal and myelin geometries. It is noteworthy that the geometry of axons and myelin does not only control AP propagation speed (Rushton, 1951; Deutsch, 1969) but also their temporal jitter (Kim et al., 2013). It seems difficult to imagine that neurons have fine-tuned their dendritic computations in a cell-type specific manner (Stuart et al., 2016), but that axons would, on the contrary, be invariant and homogeneously optimizing speed by their geometry. Additionally, we know that nervous systems adapt and fine tune a plethora of properties to accomplish specific functions (Marder and Taylor, 2011). Axons indeed adjust different parameters which impact the speed of propagation of electrical signals (Seidl et al., 2010; Ford et al., 2015; Arancibia-Carcamo et al., 2017) to obtain different speeds of computation. An illustrative example of how the geometry of axons and myelin is tuned to adjust AP propagation speed, deviating from a g-ratio of 0.6, is provided by the axon properties which encode spatial location in the avian brain (Seidl et al., 2010). The longer contralateral fibers have larger-diameter axons and longer internodal distances, compensating in this manner for an otherwise slower conduction delay relative to the shorter fibers on the ipsilateral side. In this manner, axons ensure a coincident arrival of contralateral and ipsilateral signals to the synaptic terminals. Another interesting example is provided by fibers specialized in carrying information for low-frequency sounds, which show larger diameters than those carrying high-frequency sounds, but also shorter internodes. In doing so, these fibers deviate from the classical dependence of both variables established by Rushton. This specialization is proposed to ensure a proper function of the circuit (Ford et al., 2015). Moreover, internode length and node diameter show graded properties as they approach the terminal in the auditory granular bushy cell axon, implementing an efficient invasion of the axon terminal by APs (Ford et al., 2015). We have seen that a large number of structural myelination parameters act in concert to control the speed of AP propagation (Goldman and Albus, 1968). It is noteworthy to mention that myelin has introduced new degrees of freedom in the regulation of AP speed: speed depends on distance between nodes of Ranvier, on node length, node composition, and thickness of myelin (Rushton, 1951; Ford et al., 2015; Arancibia-Carcamo et al., 2017; Wu et al., 2012). These parameters can be modulated independently or in combination, thereby increasing the number of available mechanisms by which nervous systems tune the axonal propagation of electrical signals. Furthermore, dynamical changes in axons and myelin (Sampaio-Baptista et al., 2013; McKenzie et al., 2014; Sinclair et al., 2017; Fields, 2015) suggest that nervous systems dynamically adjust the properties of their fibers to achieve the specific behaviors that they control. 9 2.3.7 Active contributions to variations of action potentials along the axon at nodes and terminals Myelinated fibers are characterized by the presence of non-myelinated sections, which concentrate the machinery to generate APs. These include nodes, the heminode (last unmyelinated portion before the terminal) and synaptic terminals. Remarkably, AP shape can suffer modifications along the axon by being modulated at each one of these loci by the active channels that they express. Each active AP regeneration site along the axon can potentially generate AP variants due to their specific state and composition of ion channels and membrane potential. An example of local variations in AP shape through active mechanisms is found at presynaptic mossy fiber terminals (present in the axon of dentate gyrus granule cells), where local potassium channels modulate spike shape (Alle et al., 2011). Another example is the activation of calcium-dependent potassium channels in Purkinje cell nodes of Ranvier. These channels repolarize membranes, de-inactivating sodium channels which can then generate fast frequency spikes, and in this manner prevent AP failure at these high frequencies (Grundemann and Clark, 2015). Additionally, changes in membrane voltage in the axon induced by active mechanisms can impact the efficiency of AP propagation. For example, in the leech touch cells, adaptation in response to repetitive AP firing consists of a hyperpolarization, resulting in the blockade of AP propagation (Van Essen, 1973). We previously described how actively-generated signals can differently be passed on by two bifurcating branches, illustrating how passive properties can filter actively-generated signals (Gu et al., 1991). Remark- ably, active properties of axonal branches can prevent failures by potentiating signals in specific branches (Cho et al., 2017). Failures that would occur as a consequence of the passive filtering of electrical signals can be prevented in a branch-specific manner by the presence of the sodium channel subunit NavβII, which potentiates AP propagation. 2.3.8 The geometries of axons and myelin are plastic Geometrical properties of both axons and myelin have been shown to be highly plastic. Indeed, the diameter of axons can change as a function of neuronal activity (Sinclair et al., 2017; Ch´ereau et al., 2017). The understanding of the mechanisms of myelination has revealed that properties of nodes and internodes are subject to plasticity (Young et al., 2013), reviewed in (Kaller et al., 2017). Interestingly, major macroscopic changes in myelination occur in response to training paradigms in mice, allowing the acquisition of motor skills (Sampaio-Baptista et al., 2013; McKenzie et al., 2014; Xiao et al., 2016), and models of injury induce strong remodelling of myelination patterns in the auditory brainstem (Sinclair et al., 2017). Changes in myelination can involve several mechanisms at time scales of hours to days: changes in internode length (Etxeberria et al., 2016), in myelin thickness (Sinclair et al., 2017) and in node length (suggested in (Arancibia-Carcamo et al., 2017), (Fig.5). 2.3.9 Nanostructures with unknown function The evolutionary appearance of myelin has led to new specialized microstructures or microdomains (namely nodes, paranodes, juxtaparanodes, heminodes). Electron microscopy and recently superresolution imaging have revealed additional structures whose contribution to axonal computation remains mysterious. These are organized at the nanometer scale in a highly regular spatial organization (D'Este et al., 2017), rising the question of the function of these periodic structures. For example, d'Este et al. show that the voltage-dependent potassium channel subunit Kv1.2 channels, found at the juxtaparanodes, correlates in space with the underlying actin cytoskeleton. An additional structure that has attracted our attention is the Schmidt-Lanterman incisure, a spiral cytoplasmic expansion from the outer tongue of myelin to the inner tongue. Incisures express gap junctions and their contribution to the electrical properties of myelin remain mysterious (Kamasawa et al., 2005). Likewise, exquisite arrangements of structures apposing glial and neuronal membranes and also with other glial membranes in the form of rosettes formed by ion channels at the paranode are still poorly understood (Rash et al., 2016). These findings open up the following question: are these highly-ordered structures an epiphenomenon resulting from the necessary anchoring to regular cytoskeletal structures, or is this spatial regularity impacting function? One would expect that further perturbing and studying those nanoscale structures in the future will further our understanding of how they contribute to axonal computations. 2.4 AXONS ARE NOT ALONE: INTRACELLULAR AND INTERCELLULAR COUPLING 2.4.1 Crosstalk of axons with other compartments Although we have done a treatment of the axon as an isolated cable as has classically been performed in the early days (Lillie, 1925; Huxley and Stampfli, 1949; Hodgkin and Huxley, 1952), the axonal cable is coupled to the somatic and dendritic compartment at one end and, additionally, directly or indirectly to electrically-coupled cells, which also functionally act like an electrical compartment (Furshpan and Potter, 1959; Alcami and Marty, 2013; Eyal et al., 2014). The electrical coupling between the axon and the soma, and indirectly to dendrites and dendritically-coupled cells, influences AP generation in the axon (Bekkers and Hausser, 2007; Eyal et al., 10 2014; Amsalem et al., 2016; Alcami, 2018; Goldwyn et al., 2019). These compartments create parallel pathways for current flow that compete with the flow of currrent in the axial resistance of the axon. Loading of these non-axonal compartments was shown to impact AP threshold and speed (Bekkers and Hausser, 2007; Eyal et al., 2014; Amsalem et al., 2016). Their contribution to the effective membrane time constant is substantial. The somato-dendritic compartment has been shown to, by this mechanism, modify both the threshold and the initial rise of the AP (Eyal et al., 2014) . It is further interesting to consider the axon in the context of synaptic integration, that is, the computation of information received at synapses. Similar to the somato-dendritic and junctional compartments, which act as current sinks, influencing the effective kinetics and strength of excitatory inputs recorded at the soma, before they reach the AIS (Norenberg et al., 2010; Alcami, 2018), the axonal membrane also behaves as a current sink, leaking current generated at synapses in the somatic and dendritic compartments with its charging time constant. This phenomenon has been shown to, through a passive mechanism, accelerate the time-course of somatically-recorded excitatory synaptic events (Mejia-Gervacio et al., 2007). It has additionally been suggested to also accelerate the time-course of spikelets generated by electrically-coupled cells (Alcami and Marty, 2013). Mejia-Gervacio and collaborators (Mejia-Gervacio et al., 2007) show that the capacitive loading of axons introduces in cerebellar molecular layer interneurons a computational time constant of about 3 ms, which is one order of magnitude slower than the faster time constant to load the somatodendritic compartment. This relatively slowly-charging process of axons accelerates the decay of excitatory postsynaptic potentials, reducing the time window for AP generation. Therefore, electrical signals that arrive to the axon are not only passively influenced by other compartments, but they also influence passive computation by the non-axonal compartments, as part of a system formed by coupled compartments. Let us now turn our attention onto the signals propagating in axons as cells detect events in their somas and dendrites, which is a consequence of the current flow of synaptic events into the axon. These signals propagate at long distances before their complete attenuation (contributing to the analog signaling in the axon that was previously introduced). The space constant at the hippocampal unmyelinated granule cell axon is in the range of hundreds of micrometers (Alle and Geiger, 2006) and the subthreshold propagation of voltage depolarizations has been shown to increase AP evoked release (Shu et al., 2006). Interestingly, these signals do not only travel orthodromically towards the terminals, but signals generated in the axon, e.g. by synaptic receptors present on the presynaptic membrane, also travel antidromically to the soma (Trigo et al., 2010). As a consequence, these axonal PSPs depolarize the soma and influence AP generation (de San Martin et al., 2015). 2.4.2 Direct coupling between axons Direct coupling between axons was suggested in early work, bringing up the concept that networks of axons may directly interact with each other (Katz and Schmitt, 1940). These interactions can be explained by two forms of electrical transmission (Fig. 6A): ephaptic transmission due to the generation of an electric field by an axon, affecting the excitability of a neighboring axon (Katz and Schmitt, 1940; Blot and Barbour, 2014; Han et al., 2018), and transmission mediated by gap junction-mediated electrical synapses (Schmitz et al., 2001; Furshpan and Potter, 1959; Watanabe and Grundfest, 1961; Bennett et al., 1963; Robertson et al., 1963). As a matter of fact, electrical synapses were first discovered in axons in a variety of preparations (Furshpan and Potter, 1959; Watanabe and Grundfest, 1961; Bennett et al., 1963; Robertson et al., 1963) . Axonal electrical synapses are expected to strongly affect electrical signals by adding a conductance pathway directly in the axon. Remarkably, axonal gap junctions allow inputs to arrive at the axon directly, blurring the pure output role of axons: an AP in a presynaptic axon induces a spikelet (a low-pass filtered version of the presynaptic AP) in the postsynaptic axon. Spikelets excite distal parts of the axon where the gap junctions are located, and when the spikelet depolarizes the membrane potential sufficiently, they are able to evoke APs (Chorev and Brecht, 2012). Interestingly, in vivo recordings from hippocampal principal cells in CA regions from rodents revealed two types of APs: one type was preceded by a shoulder which was identical to the rising phase of the spikelet, and the other was a full-blown AP lacking the shoulder (Epsztein et al., 2010). Spikelets correlate with AP firing from nearby cells, confirming that they are generated by electrically-coupled cells, and not by spontaneous AP firing of distal axons (Chorev and Brecht, 2012). Altogether, these studies suggest that hippocampal pyramidal cell axons are excited through electrical synapses in physiological conditions in behaving animals, as it had been previously shown in slices (Schmitz et al., 2001) . Axonal gap junctions have additionally been proposed to underlie fast synchronization of neuronal ensembles in both physiological and pathophysiological conditions (Schmitz et al., 2001; Roopun et al., 2010). Interestingly, axo-axonal gap junctions, by coupling axons, that is, two output structures, can have a strong impact on spike coordination and coding efficiency (Wang et al., 2017) . Last and not least, direct axo-axonal coupling can be mediated by chemical synapses onto axons (Fig. 6A) which are typically performed by a specific type of interneuron targeting the AIS of principal cells (Somogyi et al., 1998). It has been shown that the presynaptic activity of chandelier cells modulates the sodium channel dynamics on the axon initial segment (Inda et al., 2006). Terminals found in the neocortex and hippocampus 11 were found to exert inhibitory control over AP initiation (Dugladze et al., 2012). 2.4.3 Axons interact indirectly via glial cells Finally, let us consider the contribution of glial cells to axonal function as an indirect coupling pathway between axons (Fig. 6B). The biophysical properties of myelin are achieved by glial cells that produce myelin. Since several axons are typically contacted by a myelinating cell, this introduces de facto an indirect coupling pathway between axons. A number of transmission mechanisms has been described between axons and myelin (reviewed in ((Micu et al., 2017))), whose dynamic properties depend on signaling from neurons (Hines et al., 2015). 3 Conclusion Here, we have reviewed the biophysical nature of computations performed by the axon. We have shown that the axon is not just a cable on which electrical pulses propagate but rather a computational device that mod- ulates signaling and adds to the complexity of information processing in the brain. It should be appreciated that these computations are sophisticated enough to contribute to network level phenomena up to behavior in vivo. Relating axonal computation to behavioral phenomenology is still a nascent area but it will complement studies that have focused almost exclusively on somas, dendrites or a coarse grained view of the neurons. The study of axonal computation is hurdled by technical challenges, but there is already an emerging interest in developing technologies that will allow to electrophysiologically and structurally study axonal processes. Further complication arises when one realizes that axons do not act alone but in concert, exhibiting collective modes of computation conveyed by inter-axonal signaling. We propose that axonal biophysics are of vital importance for understanding not only how single neurons process information but also how neural networks coordinate their activity. Conflict of Interest Statement The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest. Author Contributions Both authors contributed equally to literature search and writing. Funding PA is funded by the Munich Center for Neurosciences. AEH is funded by the Howard Hughes Medical Institute. Acknowledgments We are grateful to Conny Kopp-Scheinflug for her feedback on previous versions of this manuscript. References Abbott, B., Howarth, J., and Ritchie, J. (1965). The initial heat production associated with the nerve impulse in crustacean and mammalian non-myelinated nerve fibres. The Journal of physiology, 178(2):368 -- 383. Akkin, T., Joo, C., and De Boer, J. F. (2007). Depth-resolved measurement of transient structural changes during action potential propagation. Biophysical journal, 93(4):1347 -- 1353. Alcami, P. (2018). Electrical synapses enhance and accelerate interneuron recruitment in response to coincident and sequential excitation. Frontiers in cellular neuroscience, 12. Alcami, P., Franconville, R., Llano, I., and Marty, A. (2012). Measuring the firing rate of high-resistance neurons with cell-attached recording. Journal of Neuroscience, 32(9):3118 -- 3130. Alcami, P. and Marty, A. (2013). Estimating functional connectivity in an electrically coupled interneuron network. Proceedings of the National Academy of Sciences, 110(49):E4798 -- E4807. 12 Alcami, P. and Pereda, A. E. (2019). Beyond plasticity: the dynamic impact of electrical synapses on neural circuits. Nature reviews. Neuroscience. Alle, H. and Geiger, J. R. (2006). Combined analog and action potential coding in hippocampal mossy fibers. Science, 311(5765):1290 -- 1293. Alle, H., Kubota, H., and Geiger, J. R. (2011). Sparse but highly efficient kv3 outpace bkca channels in action potential repolarization at hippocampal mossy fiber boutons. Journal of Neuroscience, 31(22):8001 -- 8012. Amsalem, O., Van Geit, W., Muller, E., Markram, H., and Segev, I. (2016). From neuron biophysics to orientation selectivity in electrically coupled networks of neocortical l2/3 large basket cells. Cerebral Cortex, 26(8):3655 -- 3668. Araki, T. and Otani, T. (1955). Response of single motoneurons to direct stimulation in toad's spinal cord. Journal of neurophysiology, 18(5):472 -- 485. Arancibia-Carcamo, I. L., Ford, M. C., Cossell, L., Ishida, K., Tohyama, K., and Attwell, D. (2017). Node of ranvier length as a potential regulator of myelinated axon conduction speed. Elife, 6:e23329. Ariav, G., Polsky, A., and Schiller, J. (2003). Submillisecond precision of the input-output transformation function mediated by fast sodium dendritic spikes in basal dendrites of ca1 pyramidal neurons. Journal of Neuroscience, 23(21):7750 -- 7758. Astman, N., Gutnick, M. J., and Fleidervish, I. A. (2006). Persistent sodium current in layer 5 neocortical neurons is primarily generated in the proximal axon. Journal of Neuroscience, 26(13):3465 -- 3473. Atherton, J. F., Wokosin, D. L., Ramanathan, S., and Bevan, M. D. (2008). Autonomous initiation and propaga- tion of action potentials in neurons of the subthalamic nucleus. The Journal of physiology, 586(23):5679 -- 5700. Azouz, R. and Gray, C. M. (1999). Cellular mechanisms contributing to response variability of cortical neurons in vivo. Journal of Neuroscience, 19(6):2209 -- 2223. Azouz, R., Jensen, M. S., and Yaari, Y. (1996). Ionic basis of spike after-depolarization and burst generation in adult rat hippocampal ca1 pyramidal cells. The Journal of physiology, 492 ( Pt 1)(Pt 1):211 -- 223. Baccus, S. A. (1998). Synaptic facilitation by reflected action potentials: enhancement of transmission when nerve impulses reverse direction at axon branch points. Proceedings of the National Academy of Sciences, 95(14):8345 -- 8350. Baccus, S. A., Burrell, B. D., Sahley, C. L., and Muller, K. J. (2000). Action potential reflection and failure at axon branch points cause stepwise changes in epsps in a neuron essential for learning. Journal of Neurophysiology, 83(3):1693 -- 1700. Bakiri, Y., K´arad´ottir, R., Cossell, L., and Attwell, D. (2011). Morphological and electrical properties of oligoden- drocytes in the white matter of the corpus callosum and cerebellum. The Journal of physiology, 589(3):559 -- 573. Baranauskas, G. and Martina, M. (2006). Sodium currents activate without a hodgkin and huxley-type delay in central mammalian neurons. Journal of Neuroscience, 26(2):671 -- 684. Batabyal, S., Satpathy, S., Bui, L., Kim, Y.-T., Mohanty, S., Bachoo, R., and Dav´e, D. P. (2017). Label-free optical detection of action potential in mammalian neurons. Biomedical optics express, 8(8):3700 -- 3713. Bean, B. P. (2007). The action potential in mammalian central neurons. Nature Reviews Neuroscience, 8(6):451. Bekkers, J. M. and Hausser, M. (2007). Targeted dendrotomy reveals active and passive contributions of the dendritic tree to synaptic integration and neuronal output. Proceedings of the National Academy of Sciences, 104(27):11447 -- 11452. Bennett, M. V., Aljure, E., Nakajima, Y., and Pappas, G. D. (1963). Electrotonic junctions between teleost spinal neurons: electrophysiology and ultrastructure. Science, 141(3577):262 -- 264. Bi, G.-q. and Poo, M.-m. (1998). Synaptic modifications in cultured hippocampal neurons: dependence on spike timing, synaptic strength, and postsynaptic cell type. Journal of neuroscience, 18(24):10464 -- 10472. Blot, A. and Barbour, B. (2014). Ultra-rapid axon-axon ephaptic inhibition of cerebellar purkinje cells by the pinceau. Nature neuroscience, 17(2):289. 13 Boiko, T., Van Wart, A., Caldwell, J. H., Levinson, S. R., Trimmer, J. S., and Matthews, G. (2003). Functional specialization of the axon initial segment by isoform-specific sodium channel targeting. Journal of Neuroscience, 23(6):2306 -- 2313. Borst, A. and Haag, J. (1996). The intrinsic electrophysiological characteristics of fly lobula plate tangential cells: I. passive membrane properties. Journal of computational neuroscience, 3(4):313 -- 336. Borst, J. G. G. and Sakmann, B. (1999). Effect of changes in action potential shape on calcium currents and transmitter release in a calyx -- type synapse of the rat auditory brainstem. Philosophical Transactions of the Royal Society of London. Series B: Biological Sciences, 354(1381):347 -- 355. Brachet, A., Leterrier, C., Irondelle, M., Fache, M.-P., Racine, V., Sibarita, J.-B., Choquet, D., and Dargent, B. (2010). Ankyrin g restricts ion channel diffusion at the axonal initial segment before the establishment of the diffusion barrier. The Journal of cell biology, 191(2):383 -- 395. Brette, R. (2013). Sharpness of spike initiation in neurons explained by compartmentalization. PLOS Computa- tional Biology, 9(12):1 -- 10. Calabrese, R. L. and Kennedy, D. (1974). Multiple sites of spike initiation in a single dendritic system. Brain Research, 82(2):316 -- 321. Carpenter, D. O., Hovey, M. M., and Bak, A. F. (1975). Resistivity of axoplasm. ii. internal resistivity of giant axons of squid and myxicola. The Journal of general physiology, 66(2):139 -- 148. Castelfranco, A. M. and Hartline, D. K. (2015). The evolution of vertebrate and invertebrate myelin: a theoretical computational study. Journal of computational neuroscience, 38(3):521 -- 538. Chen, W. R., Shen, G. Y., Shepherd, G. M., Hines, M. L., and Midtgaard, J. (2002). Multiple modes of action potential initiation and propagation in mitral cell primary dendrite. Journal of Neurophysiology, 88(5):2755 -- 2764. Ch´ereau, R., Saraceno, G. E., Angibaud, J., Cattaert, D., and Nagerl, U. V. (2017). Superresolution imaging reveals activity-dependent plasticity of axon morphology linked to changes in action potential conduction velocity. Proceedings of the National Academy of Sciences, 114(6):1401 -- 1406. Cho, I. H., Panzera, L. C., Chin, M., and Hoppa, M. B. (2017). Sodium channel β2 subunits prevent action potential propagation failures at axonal branch points. Journal of Neuroscience, 37(39):9519 -- 9533. Chorev, E. and Brecht, M. (2012). In vivo dual intra-and extracellular recordings suggest bidirectional coupling between ca1 pyramidal neurons. Journal of neurophysiology, 108(6):1584 -- 1593. Clark, B. A., Monsivais, P., Branco, T., London, M., and Hausser, M. (2005). The site of action potential initiation in cerebellar purkinje neurons. Nature Neuroscience, 8(2):137 -- 139. Cohen, L., Hille, B., and Keynes, R. (1970). Changes in axon birefringence during the action potential. The Journal of physiology, 211(2):495 -- 515. Colbert, C. M. and Pan, E. (2002). Ion channel properties underlying axonal action potential initiation in pyramidal neurons. Nature Neuroscience, 5(6):533 -- 538. Cole, K. (1975). Resistivity of axoplasm. i. resistivity of extruded squid axoplasm. The Journal of general physiology, 66(2):133 -- 138. Coombs, J., Curtis, D., and Eccles, J. (1957). The generation of impulses in motoneurones. The Journal of physiology, 139(2):232 -- 249. Davis, J. Q., Lambert, S., and Bennett, V. (1996). Molecular composition of the node of ranvier: identification of ankyrin-binding cell adhesion molecules neurofascin (mucin+/third fniii domain-) and nrcam at nodal axon segments. The Journal of cell biology, 135(5):1355 -- 1367. Debanne, D., Guerineau, N. C., Gahwiler, B. H., and Thompson, S. M. (1997). Action-potential propagation gated by an axonal i a-like k+ conductance in hippocampus. Nature, 389(6648):286. Debanne, D., Inglebert, Y., and Russier, M. (2019). Plasticity of intrinsic neuronal excitability. Current Opinion in Neurobiology, 54:73 -- 82. Dekker, J. P. and Yellen, G. (2006). Cooperative gating between single hcn pacemaker channels. The Journal of general physiology, 128(5):561 -- 567. 14 D'Este, E., Kamin, D., Balzarotti, F., and Hell, S. W. (2017). Ultrastructural anatomy of nodes of ranvier in the peripheral nervous system as revealed by sted microscopy. Proceedings of the National Academy of Sciences, 114(2):E191. Deutsch, S. (1969). The maximization of nerve conduction velocity. IEEE Transactions on Systems Science and Cybernetics, 5(1):86 -- 91. Doiron, B., Laing, C., Longtin, A., and Maler, L. (2002). Ghostbursting: A novel neuronal burst mechanism. Journal of Computational Neuroscience, 12(1):5 -- 25. Duflocq, A., Chareyre, F., Giovannini, M., Couraud, F., and Davenne, M. (2011). Characterization of the axon initial segment (ais) of motor neurons and identification of a para-ais and a juxtapara-ais, organized by protein 4.1 b. BMC biology, 9(1):66. Dugladze, T., Schmitz, D., Whittington, M. A., Vida, I., and Gloveli, T. (2012). Segregation of axonal and somatic activity during fast network oscillations. Science, 336(6087):1458 -- 1461. DEste, E., Kamin, D., Balzarotti, F., and Hell, S. W. (2017). Ultrastructural anatomy of nodes of ranvier in the peripheral nervous system as revealed by sted microscopy. Proceedings of the National Academy of Sciences, 114(2):E191 -- E199. El Hady, A. and Machta, B. B. (2015). Mechanical surface waves accompany action potential propagation. Nature communications, 6:6697. Epsztein, J., Lee, A. K., Chorev, E., and Brecht, M. (2010). Impact of spikelets on hippocampal ca1 pyramidal cell activity during spatial exploration. Science, 327(5964):474 -- 477. Etxeberria, A., Hokanson, K. C., Dao, D. Q., Mayoral, S. R., Mei, F., Redmond, S. A., Ullian, E. M., and Chan, J. R. (2016). Dynamic modulation of myelination in response to visual stimuli alters optic nerve conduction velocity. Journal of Neuroscience, 36(26):6937 -- 6948. Eyal, G., Mansvelder, H. D., de Kock, C. P., and Segev, I. (2014). Dendrites impact the encoding capabilities of the axon. Journal of Neuroscience, 34(24):8063 -- 8071. Fatt, P. (1957). Electric potentials occurring around a neurone during its antidromic activation. Journal of neurophysiology, 20(1):27 -- 60. Fields, R. D. (2015). A new mechanism of nervous system plasticity: activity-dependent myelination. Nature Reviews Neuroscience, 16:756 EP -- . Fleidervish, I. A., Lasser-Ross, N., Gutnick, M. J., and Ross, W. N. (2010). Na+ imaging reveals little difference in action potential -- evoked na+ influx between axon and soma. Nature neuroscience, 13(7):852. Ford, M. C., Alexandrova, O., Cossell, L., Stange-Marten, A., Sinclair, J., Kopp-Scheinpflug, C., Pecka, M., Attwell, D., and Grothe, B. (2015). Tuning of ranvier node and internode properties in myelinated axons to adjust action potential timing. Nature Communications, 6:8073 EP -- . Fourcaud-Trocm´e, N., Hansel, D., Van Vreeswijk, C., and Brunel, N. (2003). How spike generation mechanisms determine the neuronal response to fluctuating inputs. Journal of Neuroscience, 23(37):11628 -- 11640. Foust, A., Popovic, M., Zecevic, D., and McCormick, D. A. (2010). Action potentials initiate in the axon initial segment and propagate through axon collaterals reliably in cerebellar purkinje neurons. Journal of Neuroscience, 30(20):6891 -- 6902. Furshpan, E. J. and Potter, D. D. (1959). Transmission at the giant motor synapses of the crayfish. The Journal of physiology, 145(2):289 -- 325. Gasparini, S., Migliore, M., and Magee, J. C. (2004). On the initiation and propagation of dendritic spikes in ca1 pyramidal neurons. Journal of Neuroscience, 24(49):11046 -- 11056. Gentet, L. J., Stuart, G. J., and Clements, J. D. (2000). Direct measurement of specific membrane capacitance in neurons. Biophysical Journal, 79(1):314 -- 320. Golding, N. L. and Spruston, N. (1998). Dendritic sodium spikes are variable triggers of axonal action potentials in hippocampal ca1 pyramidal neurons. Neuron, 21(5):1189 -- 1200. Goldman, L. and Albus, J. S. (1968). Computation of impulse conduction in myelinated fibers; theoretical basis of the velocity-diameter relation. Biophysical Journal, 8(5):596 -- 607. 15 Goldstein, S. S. and Rall, W. (1974). Changes of action potential shape and velocity for changing core conductor geometry. Biophysical Journal, 14(10):731 -- 757. Goldwyn, J. H., Remme, M. W., and Rinzel, J. (2019). Soma-axon coupling configurations that enhance neuronal coincidence detection. PLoS computational biology, 15(3):e1006476. Grossman, Y., Parnas, I., and Spira, M. E. (1979). Differential conduction block in branches of a bifurcating axon. The Journal of physiology, 295:283 -- 305. Grundemann, J. and Clark, B. A. (2015). Calcium-activated potassium channels at nodes of ranvier secure axonal spike propagation. Cell Reports, 12(11):1715 -- 1722. Gu, X. N., Muller, K. J., and Young, S. R. (1991). Synaptic integration at a sensory-motor reflex in the leech. The Journal of physiology, 441:733 -- 754. Han, K.-S., Guo, C., Chen, C. H., Witter, L., Osorno, T., and Regehr, W. G. (2018). Ephaptic coupling promotes synchronous firing of cerebellar purkinje cells. Neuron, 100(3):564 -- 578.e3. Harris, J. J. and Attwell, D. (2012). The energetics of cns white matter. Journal of Neuroscience, 32(1):356 -- 371. Higgs, M. H., Slee, S. J., and Spain, W. J. (2006). Diversity of gain modulation by noise in neocortical neurons: regulation by the slow afterhyperpolarization conductance. Journal of Neuroscience, 26(34):8787 -- 8799. Higgs, M. H. and Spain, W. J. (2009). Conditional bursting enhances resonant firing in neocortical layer 2 -- 3 pyramidal neurons. Journal of Neuroscience, 29(5):1285 -- 1299. Hill, B. C., Schubert, E. D., Nokes, M. A., and Michelson, R. P. (1977). Laser interferometer measurement of changes in crayfish axon diameter concurrent with action potential. Science, 196(4288):426 -- 428. Hines, J. H., Ravanelli, A. M., Schwindt, R., Scott, E. K., and Appel, B. (2015). Neuronal activity biases axon selection for myelination in vivo. Nature Neuroscience, 18:683 EP -- . Hodgkin, A. L. and Huxley, A. F. (1952). A quantitative description of membrane current and its application to conduction and excitation in nerve. The Journal of physiology, 117(4):500 -- 544. Hodgkin, A. L. and Katz, B. (1949). The effect of sodium ions on the electrical activity of the giant axon of the squid. The Journal of physiology, 108(1):37 -- 77. Horn, R. and Miller, J. J. (1978). Calcium-dependent increase in spike duration during repetitive firing of aplysia axon in the presence of tea. Journal of Neurobiology, 9(5):341 -- 352. Howarth, J. (1975). Heat production in non-myelinated nerves. Philosophical Transactions of the Royal Society of London. B, Biological Sciences, 270(908):425 -- 432. Howarth, J., Keynes, R., Ritchie, J., and Von Muralt, A. (1975). The heat production associated with the passage of a single impulse in pike olfactory nerve fibres. The Journal of physiology, 249(2):349 -- 368. Hu, W., Tian, C., Li, T., Yang, M., Hou, H., and Shu, Y. (2009). Distinct contributions of nav1.6 and nav1.2 in action potential initiation and backpropagation. Nature Neuroscience, 12:996 EP -- . Huang, M., Volgushev, M., and Wolf, F. (2012). A small fraction of strongly cooperative sodium channels boosts neuronal encoding of high frequencies. PLoS One, 7(5):e37629. Huxley, A. F. and Stampfli, R. (1949). Evidence for saltatory conduction in peripheral myelinated nerve fibres. The Journal of physiology, 108(3):315 -- 339. Inda, M. C., DeFelipe, J., and Munoz, A. (2006). Voltage-gated ion channels in the axon initial segment of human cortical pyramidal cells and their relationship with chandelier cells. Proceedings of the National Academy of Sciences, 103(8):2920 -- 2925. Iwasa, K. and Tasaki, I. (1980). Mechanical changes in squid giant axons associated with production of action potentials. Biochemical and biophysical research communications, 95(3):1328 -- 1331. Izhikevich, E. M., Gally, J. A., and Edelman, G. M. (2004). Spike-timing Dynamics of Neuronal Groups. Cerebral Cortex, 14(8):933 -- 944. Kaller, M. S., Lazari, A., Blanco-Duque, C., Sampaio-Baptista, C., and Johansen-Berg, H. (2017). Myelin plasticity and behaviour -- connecting the dots. Current Opinion in Neurobiology, 47:86 -- 92. 16 Kamasawa, N., Sik, A., Morita, M., Yasumura, T., Davidson, K., Nagy, J., and Rash, J. (2005). Connexin-47 and connexin-32 in gap junctions of oligodendrocyte somata, myelin sheaths, paranodal loops and schmidt- lanterman incisures: Implications for ionic homeostasis and potassium siphoning. Neuroscience, 136(1):65 -- 86. Katz, B. and Schmitt, O. H. (1940). Electric interaction between two adjacent nerve fibres. The Journal of physiology, 97(4):471 -- 488. Keynes, R. (1951). The ionic movements during nervous activity. The Journal of physiology, 114(1-2):119 -- 150. Khaliq, Z. M. and Raman, I. M. (2006). Relative contributions of axonal and somatic na channels to action potential initiation in cerebellar purkinje neurons. Journal of Neuroscience, 26(7):1935 -- 1944. Kim, G. E., Kronengold, J., Barcia, G., Quraishi, I. H., Martin, H. C., Blair, E., Taylor, J. C., Dulac, O., Colleaux, L., Nabbout, R., and Kaczmarek, L. K. (2014). Human slack potassium channel mutations increase positive cooperativity between individual channels. Cell Reports, 9(5):1661 -- 1672. Kole, M. H., Letzkus, J. J., and Stuart, G. J. (2007). Axon initial segment kv1 channels control axonal action potential waveform and synaptic efficacy. Neuron, 55(4):633 -- 647. Kole, M. H. P. (2011). First node of ranvier facilitates high-frequency burst encoding. Neuron, 71(4):671 -- 682. Kole, M. H. P., Ilschner, S. U., Kampa, B. M., Williams, S. R., Ruben, P. C., and Stuart, G. J. (2008). Action potential generation requires a high sodium channel density in the axon initial segment. Nature Neuroscience, 11:178 EP -- . Kopp-Scheinpflug, C. and Tempel, B. L. (2015). Decreased temporal precision of neuronal signaling as a candidate mechanism of auditory processing disorder. Hearing Research, 330:213 -- 220. Kordeli, E., Lambert, S., and Bennett, V. (1995). Ankyrin a new ankyrin gene with neural-specific isoforms localized at the axonal initial segment and node of ranvier. Journal of Biological Chemistry, 270(5):2352 -- 2359. Krnjevic, K. and Miledi, R. (1959). Presynaptic failure of neuromuscular propagation in rats. The Journal of physiology, 149(1):1 -- 22. Kuba, H., Adachi, R., and Ohmori, H. (2014). Activity-dependent and activity-independent development of the axon initial segment. Journal of Neuroscience, 34(9):3443 -- 3453. Larkum, M. E., Zhu, J. J., and Sakmann, B. (1999). A new cellular mechanism for coupling inputs arriving at different cortical layers. Nature, 398(6725):338 -- 341. Larkum, M. E., Zhu, J. J., and Sakmann, B. (2001). Dendritic mechanisms underlying the coupling of the dendritic with the axonal action potential initiation zone of adult rat layer 5 pyramidal neurons. The Journal of physiology, 533(Pt 2):447 -- 466. Lazarov, E., Dannemeyer, M., Feulner, B., Enderlein, J., Gutnick, M. J., Wolf, F., and Neef, A. (2018). An axon initial segment is required for temporal precision in action potential encoding by neuronal populations. Science advances, 4(11):eaau8621. Lehnert, S., Ford, M. C., Alexandrova, O., Hellmundt, F., Felmy, F., Grothe, B., and Leibold, C. (2014). Action potential generation in an anatomically constrained model of medial superior olive axons. Journal of Neuroscience, 34(15):5370 -- 5384. Lien, C.-C. and Jonas, P. (2003). Kv3 potassium conductance is necessary and kinetically optimized for high- frequency action potential generation in hippocampal interneurons. Journal of Neuroscience, 23(6):2058 -- 2068. Lillie, R. S. (1925). Factors affecting transmission and recovery in the passive iron nerve model. The Journal of general physiology, 7(4):473 -- 507. Lorincz, A. and Nusser, Z. (2010). Molecular identity of dendritic voltage-gated sodium channels. Science, 328(5980):906 -- 909. Mackie, G. O. and Meech, R. W. (1985). Separate sodium and calcium spikes in the same axon. Nature, 313(6005):791 -- 793. Mainen, Z. F. and Sejnowski, T. J. (1996). Influence of dendritic structure on firing pattern in model neocortical neurons. Nature, 382(6589):363 -- 366. 17 Manor, Y., Koch, C., and Segev, I. (1991). Effect of geometrical irregularities on propagation delay in axonal trees. Biophysical Journal, 60(6):1424 -- 1437. Mar, A. and Drapeau, P. (1996). Modulation of conduction block in leech mechanosensory neurons. Journal of Neuroscience, 16(14):4335 -- 4343. Maranto, A. R. and Calabrese, R. L. (1984). Neural control of the hearts in the leech,hirudo medicinalis. Journal of Comparative Physiology A, 154(3):381 -- 391. Maratou, E. and Theophilidis, G. (2000). An axon pacemaker: diversity in the mechanism of generation and conduction of action potentials in snail neurons. Neuroscience, 96(1):1 -- 2. Marder, E. and Goaillard, J.-M. (2006). Variability, compensation and homeostasis in neuron and network function. Nature Reviews Neuroscience, 7(7):563. Marder, E. and Taylor, A. L. (2011). Multiple models to capture the variability in biological neurons and networks. Nature Neuroscience, 14:133 EP -- . Marx, S. O., Gaburjakova, J., Gaburjakova, M., Henrikson, C., Ondrias, K., and Marks, A. R. (2001). Coupled gating between cardiac calcium release channels (ryanodine receptors). Circulation research, 88(11):1151 -- 1158. McCormick, D. A., Shu, Y., and Yu, Y. (2007). Hodgkin and huxley model -- still standing? Nature, 445:E1 EP -- . McKenzie, I. A., Ohayon, D., Li, H., Paes de Faria, J., Emery, B., Tohyama, K., and Richardson, W. D. (2014). Motor skill learning requires active central myelination. Science, 346(6207):318 -- 322. Meeks, J. P. and Mennerick, S. (2007). Action potential initiation and propagation in ca3 pyramidal axons. Journal of Neurophysiology, 97(5):3460 -- 3472. Mejia-Gervacio, S., Collin, T., Pouzat, C., Tan, Y. P., Llano, I., and Marty, A. (2007). Axonal speeding: Shaping synaptic potentials in small neurons by the axonal membrane compartment. Neuron, 53(6):843 -- 855. Meyrand, P., Weimann, J., and Marder, E. (1992). Multiple axonal spike initiation zones in a motor neuron: serotonin activation. Journal of Neuroscience, 12(7):2803 -- 2812. Micu, I., Plemel, J. R., Caprariello, A. V., Nave, K.-A., and Stys, P. K. (2017). Axo-myelinic neurotransmission: a novel mode of cell signalling in the central nervous system. Nature Reviews Neuroscience, 19:49 EP -- . Milescu, L. S., Yamanishi, T., Ptak, K., and Smith, J. C. (2010). Kinetic properties and functional dynamics of sodium channels during repetitive spiking in a slow pacemaker neuron. Journal of Neuroscience, 30(36):12113 -- 12127. Monsivais, P., Clark, B. A., Roth, A., and Hausser, M. (2005). Determinants of action potential propagation in cerebellar purkinje cell axons. Journal of Neuroscience, 25(2):464 -- 472. Mulloney, B. and Selverston, A. (1972). Antidromic action potentials fail to demonstrate known interactions between neurons. Science, 177(4043):69 -- 72. Naundorf, B., Wolf, F., and Volgushev, M. (2006). Unique features of action potential initiation in cortical neurons. Nature, 440(7087):1060. Neef, A., Hady, A. E., Nagpal, J., Broking, K., Afshar, G., Schluter, O. M., Geisel, T., Bamberg, E., Fleischmann, R., Stuhmer, W., et al. (2013). Continuous dynamic photostimulation-inducing in-vivo-like fluctuating con- ductances with channelrhodopsins. arXiv preprint arXiv:1305.7125. Norenberg, A., Hu, H., Vida, I., Bartos, M., and Jonas, P. (2010). Distinct nonuniform cable properties optimize rapid and efficient activation of fast-spiking gabaergic interneurons. Proceedings of the National Academy of Sciences, 107(2):894. Ofer, N., Shefi, O., and Yaari, G. (2019). Axonal tree morphology and signal propagation dynamics improve interneuron classification. bioRxiv. Oh, S., Fang-Yen, C., Choi, W., Yaqoob, Z., Fu, D., Park, Y., Dassari, R. R., and Feld, M. S. (2012). Label-free imaging of membrane potential using membrane electromotility. Biophysical journal, 103(1):11 -- 18. Palay, S. L. and Chan-Palay, V. (2012). Cerebellar cortex: cytology and organization. Springer Science & Business Media. 18 Palmer, L. M., Clark, B. A., Grndemann, J., Roth, A., Stuart, G. J., and Husser, M. (2010). Rapid report: Initiation of simple and complex spikes in cerebellar purkinje cells. The Journal of Physiology, 588(10):1709 -- 1717. Palmer, L. M. and Stuart, G. J. (2006). Site of action potential initiation in layer 5 pyramidal neurons. Journal of Neuroscience, 26(6):1854 -- 1863. Perge, J. A., Koch, K., Miller, R., Sterling, P., and Balasubramanian, V. (2009). How the optic nerve allocates space, energy capacity, and information. Journal of Neuroscience, 29(24):7917 -- 7928. Pinsky, P. F. and Rinzel, J. (1994). Intrinsic and network rhythmogenesis in a reduced traub model for ca3 neurons. Journal of Computational Neuroscience, 1(1):39 -- 60. Powers, R. K., Sawczuk, A., Musick, J. R., and Binder, M. D. (1999). Multiple mechanisms of spike-frequency adaptation in motoneurones. Journal of Physiology-Paris, 93(1-2):101 -- 114. Ranade, S. S., Syeda, R., and Patapoutian, A. (2015). Mechanically activated ion channels. Neuron, 87(6):1162 -- 1179. Rasband, M. N. (2010). The axon initial segment and the maintenance of neuronal polarity. Nature Reviews Neuroscience, 11(8):552. Rash, J. E., Vanderpool, K. G., Yasumura, T., Hickman, J., Beatty, J. T., and Nagy, J. I. (2016). Kv1 channels identified in rodent myelinated axons, linked to cx29 in innermost myelin: support for electrically active myelin in mammalian saltatory conduction. Journal of Neurophysiology, 115(4):1836 -- 1859. Rios, J. C., Rubin, M., Martin, M. S., Downey, R. T., Einheber, S., Rosenbluth, J., Levinson, S. R., Bhat, M., and Salzer, J. L. (2003). Paranodal interactions regulate expression of sodium channel subtypes and provide a diffusion barrier for the node of ranvier. Journal of Neuroscience, 23(18):7001 -- 7011. Ritchie, J. M. and Rogart, R. B. (1977). Density of sodium channels in mammalian myelinated nerve fibers and nature of the axonal membrane under the myelin sheath. Proceedings of the National Academy of Sciences of the United States of America, 74(1):211 -- 215. Robertson, J. D., Bodenheimer, T. S., and Stage, D. E. (1963). The ultrastructure of mauthner cell synapses and nodes in goldish brains. The Journal of cell biology, 19(1):159 -- 199. Roopun, A. K., Simonotto, J. D., Pierce, M. L., Jenkins, A., Nicholson, C., Schofield, I. S., Whittaker, R. G., Kaiser, M., Whittington, M. A., Traub, R. D., and Cunningham, M. O. (2010). A nonsynaptic mechanism underlying interictal discharges in human epileptic neocortex. Proceedings of the National Academy of Sciences, 107(1):338 -- 343. Rosenbluth, J. (1976). Intramembranous particle distribution at the node of ranvier and adjacent axolemma in myelinated axons of the frog brain. Journal of Neurocytology, 5(6):731 -- 745. Rudolph, M. and Destexhe, A. (2003). Characterization of subthreshold voltage fluctuations in neuronal mem- branes. Neural Computation, 15(11):2577 -- 2618. Rush, A. M., Dib-Hajj, S. D., and Waxman, S. G. (2005). Electrophysiological properties of two axonal sodium channels, nav1.2 and nav1.6, expressed in mouse spinal sensory neurones. The Journal of physiology, 564(Pt 3):803 -- 815. Rushton, W. A. H. (1951). A theory of the effects of fibre size in medullated nerve. The Journal of physiology, 115(1):101 -- 122. Sampaio-Baptista, C., Khrapitchev, A. A., Foxley, S., Schlagheck, T., Scholz, J., Jbabdi, S., DeLuca, G. C., Miller, K. L., Taylor, A., Thomas, N., Kleim, J., Sibson, N. R., Bannerman, D., and Johansen-Berg, H. (2013). Motor skill learning induces changes in white matter microstructure and myelination. Journal of Neuroscience, 33(50):19499 -- 19503. Sawczuk, A., Powers, R. K., and Binder, M. D. (1997). Contribution of outward currents to spike-frequency adaptation in hypoglossal motoneurons of the rat. Journal of neurophysiology, 78(5):2246 -- 2253. Schmidt, D., del M´armol, J., and MacKinnon, R. (2012). Mechanistic basis for low threshold mechanosensitivity in voltage-dependent k+ channels. Proceedings of the National Academy of Sciences, 109(26):10352 -- 10357. Schmidt, D. and MacKinnon, R. (2008). Voltage-dependent k+ channel gating and voltage sensor toxin sensitivity depend on the mechanical state of the lipid membrane. Proceedings of the National Academy of Sciences, 105(49):19276 -- 19281. 19 Schmidt-Hieber, C., Jonas, P., and Bischofberger, J. (2008). Action potential initiation and propagation in hippocampal mossy fibre axons. The Journal of Physiology, 586(7):1849 -- 1857. Schmitz, D., Schuchmann, S., Fisahn, A., Draguhn, A., Buhl, E. H., Petrasch-Parwez, E., Dermietzel, R., Heinemann, U., and Traub, R. D. (2001). Axo-axonal coupling: A novel mechanism for ultrafast neuronal communication. Neuron, 31(5):831 -- 840. Seidl, A. H., Rubel, E. W., and Harris, D. M. (2010). Mechanisms for adjusting interaural time differences to achieve binaural coincidence detection. Journal of Neuroscience, 30(1):70 -- 80. Sheffield, M. E. J., Best, T. K., Mensh, B. D., Kath, W. L., and Spruston, N. (2010). Slow integration leads to persistent action potential firing in distal axons of coupled interneurons. Nature Neuroscience, 14:200 EP -- . Shu, Y., Hasenstaub, A., Duque, A., Yu, Y., and McCormick, D. A. (2006). Modulation of intracortical synaptic potentials by presynaptic somatic membrane potential. Nature, 441(7094):761 -- 765. Sinclair, J. L., Fischl, M. J., Alexandrova, O., He, M., Grothe, B., Leibold, C., and Kopp-Scheinpflug, C. (2017). Sound-evoked activity influences myelination of brainstem axons in the trapezoid body. Journal of Neuroscience, 37(34):8239 -- 8255. Song, A.-h., Wang, D., Chen, G., Li, Y., Luo, J., Duan, S., and Poo, M.-m. (2009). A selective filter for cytoplasmic transport at the axon initial segment. Cell, 136(6):1148 -- 1160. Spira, M. E., Oren, R., Dormann, A., and Gitler, D. (2003). Critical calpain-dependent ultrastructural alterations underlie the transformation of an axonal segment into a growth cone after axotomy of cultured aplysia neurons. Journal of Comparative Neurology, 457(3):293 -- 312. Spirou, G., Chirila, F., von Gersdorff, H., and Manis, P. (2008). Heterogeneous ca2+ influx along the adult calyx of held: A structural and computational study. Neuroscience, 154(1):171 -- 185. Stasheff, S. F., Hines, M., and Wilson, W. A. (1993). Axon terminal hyperexcitability associated with epilepto- genesis in vitro. i. origin of ectopic spikes. Journal of Neurophysiology, 70(3):961 -- 975. Stuart, G. and Sakmann, B. (1995). Amplification of epsps by axosomatic sodium channels in neocortical pyramidal neurons. Neuron, 15(5):1065 -- 1076. Stuart, G., Spruston, N., and Hausser, M. (2016). Dendrites. Oxford University Press. Stuart, G., Spruston, N., Sakmann, B., and Hausser, M. (1997). Action potential initiation and backpropagation in neurons of the mammalian cns. Trends in Neurosciences, 20(3):125 -- 131. Sudhof, T. C. and Malenka, R. C. (2008). Understanding synapses: Past, present, and future. Neuron, 60(3):469 -- 476. Tasaki, I. and Byrne, P. M. (1992). Heat production associated with a propagated impulse in bullfrog myelinated nerve fibers. The Japanese journal of physiology, 42(5):805 -- 813. Tasaki, I. and Iwasa, K. (1982a). Further studies of rapid mechanical changes in squid giant axon associated with action potential production. The Japanese journal of physiology, 32(4):505 -- 518. Tasaki, I. and Iwasa, K. (1982b). Rapid pressure changes and surface displacements in the squid giant axon associated with production of action potentials. The Japanese journal of physiology, 32(1):69 -- 81. Tchumatchenko, T., Malyshev, A., Geisel, T., Volgushev, M., and Wolf, F. (2010). Correlations and synchrony in threshold neuron models. Physical review letters, 104(5):058102. Tchumatchenko, T., Malyshev, A., Wolf, F., and Volgushev, M. (2011). Ultrafast population encoding by cortical neurons. Journal of Neuroscience, 31(34):12171 -- 12179. Trigo, F. F., Bouhours, B., Rostaing, P., Papageorgiou, G., Corrie, J. E., Triller, A., Ogden, D., and Marty, A. (2010). Presynaptic miniature gabaergic currents in developing interneurons. Neuron, 66(2):235 -- 247. Tyler, W. J. (2012). The mechanobiology of brain function. Nature Reviews Neuroscience, 13(12):867. Van Essen, D. C. (1973). The contribution of membrane hyperpolarization to adaptation and conduction block in sensory neurones of the leech. The Journal of Physiology, 230(3):509 -- 534. Wang, S., Borst, A., Zaslavsky, N., Tishby, N., and Segev, I. (2017). Efficient encoding of motion is mediated by gap junctions in the fly visual system. PLOS Computational Biology, 13(12):1 -- 22. 20 Watanabe, A. and Grundfest, H. (1961). Impulse propagation at the septal and commissural junctions of crayfish lateral giant axons. The Journal of General Physiology, 45(2):267 -- 308. Wei, W. and Wolf, F. (2011). Spike onset dynamics and response speed in neuronal populations. Physical Review Letters, 106(8):088102. Wollner, D. A. and Catterall, W. A. (1986). Localization of sodium channels in axon hillocks and initial segments of retinal ganglion cells. Proceedings of the National Academy of Sciences, 83(21):8424 -- 8428. Wu, L. M. N., Williams, A., Delaney, A., Sherman, D. L., and Brophy, P. J. (2012). Increasing internodal distance in myelinated nerves accelerates nerve conduction to a flat maximum. Current Biology, 22(20):1957 -- 1961. Xiao, L., Ohayon, D., McKenzie, I. A., Sinclair-Wilson, A., Wright, J. L., Fudge, A. D., Emery, B., Li, H., and Richardson, W. D. (2016). Rapid production of new oligodendrocytes is required in the earliest stages of motor-skill learning. Nature Neuroscience, 19:1210 EP -- . Xu, K. and Terakawa, S. (1999). Fenestration nodes and the wide submyelinic space form the basis for the unusually fast impulse conduction of shrimp myelinated axons. Journal of Experimental Biology, 202(15):1979 -- 1989. Yamamoto, T., Samejima, A., and Oka, H. (1990). The mode of synaptic activation of pyramidal neurons in the cat primary somatosensory cortex: an intracellular hrp study. Experimental Brain Research, 80(1):12 -- 22. Yang, Y., Ramamurthy, B., Neef, A., and Xu-Friedman, M. A. (2016). Low somatic sodium conductance enhances action potential precision in time-coding auditory neurons. Journal of Neuroscience, 36(47):11999 -- 12009. Yau, K. W. (1976). Receptive fields, geometry and conduction block of sensory neurones in the central nervous system of the leech. The Journal of physiology, 263(3):513 -- 538. Young, K. M., Psachoulia, K., Tripathi, R. B., Dunn, S.-J., Cossell, L., Attwell, D., Tohyama, K., and Richardson, W. D. (2013). Oligodendrocyte dynamics in the healthy adult cns: evidence for myelin remodeling. Neuron, 77(5):873 -- 885. Yu, Y., Shu, Y., and McCormick, D. A. (2008). Cortical action potential backpropagation explains spike threshold variability and rapid-onset kinetics. Journal of Neuroscience, 28(29):7260 -- 7272. 21 Figures Figure 1: Diversity of axons: morphology and function. A. Giant fibers in the crawfish contact each other by electrical synapses. From Furshpan and Potter, 1958. B. Cerebellar basket cells collaterals contact several Purkinje cells, which they inhibit at the basket structure via chemical synapses, and the pinceau via ephaptic inhibition. From Palay and Chan-Palay, Springer Verlag Berlin, 1974. C. The climbing fiber in the cerebellum shows different axonal specializations: the glomerular collateral in the granule cell layer and the climbing fiber wrapping around Purkinje cells. From Palay and Chan-Palay, Springer Verlag Berlin, 1974. D. The honeybee axon of the BC cell projects onto the left and right hemispheres, targeting a number of regions (one of the target regions is labelled in yellow). From Zwaka et al., 2016. 22 Figure 2: Overview of axonal electrical signaling. Axonal signaling is formed by a mixture of analog and digital signaling. Insets: membrane potential traces. Action potentials can propagated orthodromically or antidromically. Right, in blue, active sites with the machinery to generate or regenerate action potentials. AIS, axon initial segment. Figure 3: Overview of Action Potential Diversity. A Ultrastructure of the axon initial segment. Green is sodium channels, orange circles are Ankryin G molecules, red lines are beta IV spectrin and orange lines are the actin rings. B1. Left, example of fast rising action potential from a neocortical pyramidal cell. Right, phase plane plot of this cell. B2. Left, example of a stellate cell from the cochlear nucleus with two components: a fast rising phase of the action potential contributed by the AIS and a second phase contributed by sodium by the somatodendritic compartment. Left, phase plane plot of this cell. 23 Figure 4: Equivalent circuits of axons and structural biophysical specializations that improve conduction Top left. An axon can be reduced to an equivalent electrical circuit involving an axial resistor (parallel to the membrane) conveyed by the intracellular medium of the axon (axoplasm), and a parallel RC circuit formed by the membrane resistance and membrane capacitance. Top right, conduction can be improved by decreasing axon diameter, and thereby longitudinal resistance to current flow, allowing for a larger current flow in the axon relative to the membrane resistor, and thereby a larger space constant and faster conduction. Bottom, an alternative circuit modification occurs with myelination: additional capacitances conveyed by myelin in series with axonal capacitance reduce the effective capacitance and additional resistance increase the effective resistance, increasing propagation speed and space constant, respectively. Note that the battery associated to each membrane resistor has been omitted for simplification purposes. Orange arrows represent the current flow, and their thickness is indicative of the relative current flow in the axoplasm and in the radial direction out of the axon. 24 Figure 5: Structure-function plasticity A. A1. Axon structural plasticity. A2. Functional implications: action potential latency, measured by antidromically stimulating axons before and after a plasticity protocol, decreases several minutes after the protocol. From Chereau et al. (2015) B. B1. Myelin structural plasticity can involve changes in internode length (Etxebarria et al., 2016), in myelin thickness (Sinclair et al., 2017) or in node length (suggested in Arancibia-Carcamo, 2017). B2. Functional implications of changes in myelination: MBP optical density, putatively proportional to the amount of myelin positively correlates in the learning rate in a motor skill learning paradigm. From Sampaio-Baptista et al. (2013). Figure 6: Coupling modalities between axons. A. Three types of synapse-mediated coupling: through axo-axonal gap junctions (left), ephaptic coupling (middle), and chemical axonic synapses (right). B. Indirect coupling via myelinating cells. 25
1910.10559
1
1910
2019-10-23T13:42:53
Working memory facilitates reward-modulated Hebbian learning in recurrent neural networks
[ "q-bio.NC", "cs.NE" ]
Reservoir computing is a powerful tool to explain how the brain learns temporal sequences, such as movements, but existing learning schemes are either biologically implausible or too inefficient to explain animal performance. We show that a network can learn complicated sequences with a reward-modulated Hebbian learning rule if the network of reservoir neurons is combined with a second network that serves as a dynamic working memory and provides a spatio-temporal backbone signal to the reservoir. In combination with the working memory, reward-modulated Hebbian learning of the readout neurons performs as well as FORCE learning, but with the advantage of a biologically plausible interpretation of both the learning rule and the learning paradigm.
q-bio.NC
q-bio
Working memory facilitates reward-modulated Hebbian learning in recurrent neural networks Roman Pogodin1,∗, Dane Corneil2, Alexander Seeholzer2, Joseph Heng2, Wulfram Gerstner2 1Gatsby Computational Neuroscience Unit, University College London, London, UK 2 School of Computer and Communication Sciences and School of Life Sciences, Brain Mind Institute, École Polytechnique Fédérale de Lausanne, Lausanne, Switzerland ∗[email protected] Abstract Reservoir computing is a powerful tool to explain how the brain learns temporal se- quences, such as movements, but existing learning schemes are either biologically implausible or too inefficient to explain animal performance. We show that a net- work can learn complicated sequences with a reward-modulated Hebbian learning rule if the network of reservoir neurons is combined with a second network that serves as a dynamic working memory and provides a spatio-temporal backbone sig- nal to the reservoir. In combination with the working memory, reward-modulated Hebbian learning of the readout neurons performs as well as FORCE learning, but with the advantage of a biologically plausible interpretation of both the learning rule and the learning paradigm. 1 Introduction Learning complex temporal sequences that extend over a few seconds -- such as a movement to grab a bottle or to write a number on the blackboard -- looks easy to us but is challenging for computational brain models. A common framework for learning temporal sequences is reservoir computing (alternatively called liquid computing or echo-state networks) [1, 2, 3]. It combines a reservoir, a recurrent network of rate units with strong, but random connections [4], with a linear readout that feeds back to the reservoir. Training of the readout weights with FORCE, a recursive least-squares estimator [1], leads to excellent performance on many tasks such as motor movements. The FORCE rule is, however, biologically implausible: update steps of synapses are rapid and large, and require an immediate and precisely timed feedback signal. A more realistic alternative to FORCE is the family of reward-modulated Hebbian learning rules [5, 6, 7], but plausibility comes at a price: when the feedback (reward minus expected reward) is given only after a long delay, reward-modulated Hebbian plasticity is not powerful enough to learn complex tasks. Here we combine the reservoir network with a second, more structured network that stores and updates a two-dimension continuous variable as a "bump" in an attractor [8, 9]. The activity of the attractor network acts as a dynamic working memory and serves as input to the reservoir network (fig. 1). Our approach is related to that of feeding an abstract oscillatory input [10] or a "temporal backbone signal" [11] into the reservoir in order to overcome structural weaknesses of reservoir computing that arise if large time spans need to be covered. In computational experiments, we show that a dynamic working memory that serves as an input to a reservoir network facilitates reward-modulated Hebbian learning in multiple ways: it makes a biologically plausible three-factor rule as efficient as FORCE; it admits a delay in the feedback signal; and it allows a single reservoir network to learn and perform multiple tasks. NeurIPS 2019 workshop "Real Neurons & Hidden Units: Future directions at the intersection of neuroscience and artificial intelligence", Vancouver, Canada. 2 Model Our architecture is simple: the attractor network (the "memory") receives some task-specific input and produces a robust two-dimensional neural trajectory; the reservoir network (the "motor cortex") shapes its dynamics with this trajectory, and produces a potentially high-dimensional output (fig. 1). Figure 1: Model architecture: a moving 2D bump (left: activity bump (red) surrounded by inactive neurons (blue)) in an attractor network (left circle with two bump trajectories) projects to a reservoir (right circle); the output z(t) is read out from the reservoir and approximates the target function. Attractor network. Following [9], the bump attractor consists of 2500 neurons evolving as τm x = −x + [Jx + e − h]+ , τa h = −h + s x , (1) where x is the vector of firing rates evolving with time constant τm, e is the task-specific external input, h is an adaptation variable with time constant τa and s is the strength of adaptation. The weight matrix J = Js + Jh has two parts. The symmetric part Js creates a two-dimensional translation- invariant structure resulting in bump-like stable activity patterns, whereas Jh represents structural noise. Due to the adaptation h, the bump moves across a path defined by the initial conditions and structural noise, creating long -- lasting reliable activity patterns which also depend on the input e. Reservoir network. The reservoir learns to approximate a target function f (t) with the output z(t) by linearly combining the firing rate r with readout weights Wro: z = Wror + ηηη ≡ z + ηηη with readout noise ηηη. We use the same number of neurons (1000) and parameters as [1, 6], τ u = −u + λWrecr + Wfbz + c Wattrx , r = tanh(u) + ξξξ , (2) where u is the membrane potential, ξξξ is the firing rate noise, Wattr scales attractor input with coupling c, Wrec and λ regulate chaotic activity [4], and Wfb implements the feedback loop. Learning rule. We use the reward-modulated Hebbian rule of [6] for the readout weights Wro, ∆Wro(t) = η(t)M (t)(z(t) − ¯z(t))r(cid:62)(t) , (3) where ¯x denotes low-pass filtering of x, such that z(t) − ¯z(t) ≈ ηηη(t). The reward modulation M (t) tracks performance P (t) as η(t) = η0/(1 + t/τη) , (cid:26)1, P (t) > ¯P (t) , 0, P (t) ≤ ¯P (t) . P (t) = −(cid:107)f (t) − z(t)(cid:107)2 , M (t) = (4) The update rule is an example of a NeoHebbian three-factor learning rule [12] and mimics gradient descent if we ignore the feedback loop [5]. For model details, see appendix A. 3 Experiments In fig. 2, the learning rules are compared on 50 target functions sampled from a Gaussian Process (GP) with exponential squared kernel (σ2 = 104 to match the complexity of hand-picked functions from [1, 6]). After each training period, we measure performance with normalized cross-correlation between the output and the target (ranging from -1 to 1, where 1 is a perfect match) on a single trial with frozen weights. Details are provided in appendix A; code: https://github.com/neuroai-workshop- anon-1224113/working-memory-facilitating-reservoir-learning. 3.1 Reward-modulated Hebbian learning with attractor input reaches FORCE performance When tested on one-second signals similar to those of [1, 6] (two insets in fig. 2), the full network with attractor input and reward-modulated Hebbian learning learns faster and more reliably than 2 reward-modulated Hebbian learning without the input from the attractor network. After about 90 training trials, the full network achieves the performance of the FORCE rule (for which training error approaches one in the first trial, [1], while test error does so after 30-50 trials, [6]; fig. 2A). For target signals that extend over 10 seconds (same smoothness of the target functions, two insets in fig. 2B), the reward-modulated Hebbian rule achieves a performance of 1 after 200 trials if combined with input from the attractor network (fig. 2B) but fails completely without the attractor network (tuning of the hyperparameters on a logarithmic scale did not help; data not shown). Thus a three- factor learning rule succeeds to learn complex tasks if combined with a temporally structured input from the attractor network. Figure 2: A. Mean test performance as a function of trials with a reward-modulated Hebbian (rmHebb) learning rule combined with (red) or without (green) input from attractor. Inset: target function over 1s and examples of the output after training. B. Same, but for a target signal that spans over 10s. Shaded area shows standard deviation. The FORCE rule achieves a performance of 1 in all trials. C. Learning performance for different couplings c of the attractor output to reservoir input (c = 1 normal coupling, c = 0 no coupling) when updates are delayed to the end of every one-second trial. 3.2 Attractor input allows realistic frequency of weight updates FORCE learning needs a feedback signal at every time step. Standard reward-modulated Hebbian learning can support very small delays, but fails if updates are less frequent than every few ms [6]. In our approach (fig. 2C), proposed updates are summed up in the background, but applied only at the end of a one-second trial. We find that even with such a temporally sparse update, learning is still possible. The input from the dynamic working memory is necessary to achieve this task: when the strength of the input from the attractor network gradually decreases, performance drops; in the total absence of attractor input (c = 0.0; note that the reservoir still receives weak input noise) learning completely fails. Strikingly, delayed updates do not hurt performance, and the system achieves high (> 0.9) cross-correlation in fewer than 100 training trials if the input from the attractor network is strong enough. The transient drop in performance shortly after the start in fig. 2C is likely due to Wro = 0 in the beginning, meaning that the output is uncorrelated with the firing rates, and therefore the cumulative weight update does not approximate gradient information. 3.3 Working memory translates to efficient reservoir learning of multiple signals It is well known that reservoir networks can learn multiple tasks given different initial conditions with both FORCE [1] and the reward-modulated Hebbian rule [6]. We want to check whether this also holds for our approach. We conjecture that different inputs to the attractor network generate unique neural trajectories [9] that can be exploited by the reservoir network. To test this hypothesis, we train the network to produce hand-written digits. The static input to the attractor comes from the pre-processed MNIST dataset (network inputs are taken from one of the last layers of a deep network trained to classify MNIST) in order to provide a realistic input to the attractor network which transforms the static input into dynamic trajectories (noiseless, fig. 3B, and noisy, fig. 3D). We record 50 attractor trajectories used for training (used 4 times each, resulting in 2000 training trials) and 50 for testing of each digit (1 second each), where each trajectory corresponds to a distinct input pattern. The reservoir learns a single drawing for each class. The variance of the structural noise in the attractor network is 3 times larger compared to the previous experiments in order to produce more robust bump trajectories (fig. 3D). The reward-modulated Hebbian rule masters 10 out of 10 digits when driven by a noiseless input from the attractor network (fig. 3A). In the presence of noise in the attractor network (fig. 3D), the 3 050100150200250300350400period0.00.20.40.60.81.0cross-corr ± STDrmHebbrmHebb + attractorAtargetoutput050100150200250300350400period0.00.20.40.60.81.0cross-corr ± STD020406080100period0.00.20.40.60.81.0cross-corr ± SEMBCc=0.0c=0.1c=0.2c=0.3c=0.4c=0.5-1 performance is imperfect for "five" and "six" (fig. 3C). We checked that FORCE learning with the same noisy input did not improve the performance (data not shown). Note that a linear readout of the attractor (without the reservoir) would be insufficient: first, sometimes single digit trajectories are very dissimilar (e.g. the different zero's in fig. 3D); second, at points where trajectories cross each other, a delay-less linear readout must produce the same output, no matter what the digit is. Figure 3: A. Targets (green) and test outputs (red) for reward-modulated Hebbian rule with noiseless attractor input. B. Noiseless attractor trajectories for each digit. C. Same as A but with noisy attractor trajectories. D. Same as B but static input's noise resulted in noisy trajectories (individual lines). 4 Discussion We showed that a dynamic working memory can facilitate learning of complex tasks with biologically plausible three-factor learning rules. Our results indicate that, when combined with a bump attractor, reservoir computing with reward-modulated learning can be as efficient as FORCE [1], a widely used but biologically unrealistic rule. The proposed network relies on a limited number of trajectories in the attractor network. To increase its capacity, a possible future direction would be to combine input from the attractor network with another, also input-specific, but transient input that would bring the reservoir into a different initial state. In this case the attractor network would work as a time variable (as in [9]), and the other input as the control signal (as in [1]). Apart from the biological relevance, the proposed method might be used for real-world applications of reservoir computing (e.g. wind forecasting [13]) as it is computationally less expensive than FORCE. It might also be an interesting alternative for learning in neuromorphic devices. Acknowledgments This research was supported by the Gatsby Charitable Foundation, Swiss National Science Foundation (no. 200020 - 184615) and by the European Union Horizon 2020 Framework Program under grant agreement no. 785907 (HumanBrain Project, SGA2). The authors thank Moritz Deger for an earlier version of the reservoir code and Jorge Aurelio Menendez for useful comments on the manuscript. References [1] D. Sussillo and L.F. Abbott. Neuron, 63:544 -- 557, 2009. [2] W. Maass, T. Natschläger, and H. Markram. Neural Computation, 14:2531 -- 2560, 2002. [3] H. Jaeger and H. Haas. Science, 304:78 -- 80, 2004. [4] H. Sompolinsky, A. Crisanti, and H. J. Sommers. Physical Review Letters, 61:259 -- 262, 1988. [5] R. Legenstein, S. M. Chase, A. B. Schwartz, and W. Maass. Journal of Neuroscience, 30:8400 -- 8410, 2010. [6] G.M. Hoerzer, R. Legenstein, and W. Maass. Cerebral cortex, 24:677 -- 90, 2014. [7] N. Frémaux and W. Gerstner. Frontiers in Neural Circuits, 9:85, 2016. [8] S. Wu, K. Hamaguchi, and S. Amari. Neural computation, 20:994 -- 1025, 2008. [9] V. Itskov, C. Curto, E. Pastalkova, and G. Buzsaki. Journal of Neuroscience, 31:2828 -- 2834, 2011. [10] P. Vincent-Lamarre, G. Lajoie, and J.P. Thivierge. J. Comput. Neurosci., 41:305 -- 322, 2016. [11] W. Nicola and C. Clopath. Nat. Comm., 8:2208, 2017. [12] W. Gerstner, M. Lehmann, V. Liakoni, D. Corneil, and J. Brea. Front. Neural Circ., 12:53, 2018. [13] E. López, C. Valle, H. Allende, E. Gil, and H. Madsen. Energies, 11:526, 2018. 4 DABC A Supplementary Materials Simulation details. Both networks were simulated with the Euler method with the step size dt = 1 ms. The attractor network dynamics was recorded after a 100 ms warm up period to allow creation of the bump solution, during which it received additional input from the images in section 3.3. Training was done consequently, without breaks in the dynamics between trials. For testing, the network started from the preceding training state and continued with frozen weights. After testing, the pre-training activity was restored. The code for experiments is available at https://github.com/neuroai-workshop- anon-1224113/working-memory-facilitating-reservoir-learning. Test functions. Gaussian process test function were drawn from (5) Forcing both ends of the function to be zero and denoting x = (0, T − 1)(cid:62), z = (1, . . . , T − 2)(cid:62), we sample test functions as fGP ∼ GP (0, K) , K(x, y) = exp(cid:0)−(x − y)2/(2σ2)(cid:1) . fGP (1, . . . , T − 2) ∼ N(cid:0)0, K(z, x)K−1(x, x)K(x, z)(cid:1) , (6) where T is either 103 (short tasks) or 104 (long tasks). We chose σ2 to roughly match the complexity of targets from [6] (σ2 = 104). 50 random functions were tested on 50 random reservoirs that nevertheless received the same attractor input (Wattr was not resampled). In section 3.3, the same reservoir was used for all runs. The noisy input for section 3.3 was taken from an intermediate layer of a deep network trained to classify MNIST, and the noiseless input stimulated only a 5 by 5 square of neurons (unique for each digit). Attractor network parameters. The time constants were τm = 30 ms, τa = 400 ms. Adaptation strength was s = 1.5. The external input e was drawn independently for each neuron from a Gaussian distribution N (1, 0.00252). In section 3.3, the task-specific input was added to the noisy one. For the connectivity matrix J = Js + Jh, the noisy part was drawn independently as (Jh)ij ∼ N (0, σ2/Nattr), with Nattr = 2500 and σ = 2 in all experiments except for section 3.3, where we used σ = 6 for more robust trajectories. The symmetric part arranged the neurons on a 2D grid, such that every neuron i had its coordinates xi and yi ranging from 0 to 49. The connectivity led to mutual excitation of nearby neurons and inhibition of the distant ones, (Js)ij = −0.375 + exp(cid:0)−d(i, j)2/2(cid:1) , (7) (cid:113) 1√ 2π d(i, j) = π L (min(xi − xj, L − xi − xj))2 + (min(yi − yj, L − yi − yj))2, L = 50. (8) The bump center (used in fig. 3B and D) corresponded to the mean of the activity on the torus. Denoting activity of each neuron as r(x, y), the center on the x axis was calculated as xcenter = L 2π angle 1 L r(xn, ym)e2πin/L , L = 50, (9) where "angle" computes the counterclockwise angle of a complex variable (ranging from 0 to 2π). Reservoir network parameters. The time constant was τ = 50 ms, and total coupling strength was λ = 1.5. The readout weights Wro were initialized to zero. The feedback weights were drawn independently from a uniform distribution as (Wfb)ij ∼ U(−1, 1). Both the recurrent connections and the weights from the attractor to the reservoir were drawn independently as (Wrec)ij, (Wattr)ij ∼ N (0, 1/pNres) · Be(p), with p = 0.1, Nres = 1000, and Be being the Bernoulli distribution. A new reservoir, and thus Wrec, was sampled for each new test function. The matrix Wrec was the same for all tasks except the last one in section 3.3. State noise ξξξ and exploratory noise ηηη were generated independently from the uniform distribution as ξi ∼ U(−0.05, 0.05), ηi ∼ U(−0.5, 0.5). When attractor was present, the reservoir neurons also received weak independent noise drawn from N (0, 0.0025). Learning rule. Low-pass filtering was done as ¯x(t + dt) = ¯x(t) + dt (x(t) − ¯x(t))/τf , τf = 5 ms, ¯x(0) = 0. (10) The learning rate η(t) was computed as η(t) = η0/(1 + t/τl) (η0 = 5 · 10−4, τl = 2 · 104 ms) and held at η0 in section 3.2 to make conclusions independent of the decay. 5 (cid:32)L−1(cid:88) L−1(cid:88) n=0 m=0 (cid:33)
1901.00708
1
1901
2019-01-03T13:17:19
Decoding Hand Kinematics from Local Field Potentials Using Long Short-Term Memory (LSTM) Network
[ "q-bio.NC" ]
Local field potential (LFP) has gained increasing interest as an alternative input signal for brain-machine interfaces (BMIs) due to its informative features, long-term stability, and low frequency content. However, despite these interesting properties, LFP-based BMIs have been reported to yield low decoding performances compared to spike-based BMIs. In this paper, we propose a new decoder based on long short-term memory (LSTM) network which aims to improve the decoding performance of LFP-based BMIs. We compare offline decoding performance of the proposed LSTM decoder to a commonly used Kalman filter (KF) decoder on hand kinematics prediction tasks from multichannel LFPs. We also benchmark the performance of LFP-driven LSTM decoder against KF decoder driven by two types of spike signals: single-unit activity (SUA) and multi-unit activity (MUA). Our results show that LFP-driven LSTM decoder achieves significantly better decoding performance than LFP-, SUA-, and MUA-driven KF decoders. This suggests that LFPs coupled with LSTM decoder could provide high decoding performance, robust, and low power BMIs.
q-bio.NC
q-bio
Decoding Hand Kinematics from Local Field Potentials Using Long Short-Term Memory (LSTM) Network Nur Ahmadi1,2, Student Member, IEEE, Timothy G. Constandinou1,2, Senior Member, IEEE, and Christos-Savvas Bouganis1, Senior Member, IEEE 9 1 0 2 n a J 3 ] . C N o i b - q [ 1 v 8 0 7 0 0 . 1 0 9 1 : v i X r a Abstract -- Local field potential (LFP) has gained increas- ing interest as an alternative input signal for brain-machine interfaces (BMIs) due to its informative features, long-term stability, and low frequency content. However, despite these interesting properties, LFP-based BMIs have been reported to yield low decoding performances compared to spike-based BMIs. In this paper, we propose a new decoder based on long short-term memory (LSTM) network which aims to improve the decoding performance of LFP-based BMIs. We compare offline decoding performance of the proposed LSTM decoder to a commonly used Kalman filter (KF) decoder on hand kinematics prediction tasks from multichannel LFPs. We also benchmark the performance of LFP-driven LSTM decoder against KF decoder driven by two types of spike signals: single- unit activity (SUA) and multi-unit activity (MUA). Our results show that LFP-driven LSTM decoder achieves significantly better decoding performance than LFP-, SUA-, and MUA- driven KF decoders. This suggests that LFPs coupled with LSTM decoder could provide high decoding performance, robust, and low power BMIs. I. INTRODUCTION Brain machine interfaces (BMIs) hold the promise to restore lost motor function in persons with neurological disorders (e.g. spinal cord injury) by enabling interaction with the environment through their neural activity. A key component of a BMI system is the decoder which translates neural activity (i.e. intentions) into motor commands to control external devices such as computer cursors and robotic arms, or the subject's own muscles. To date, high decod- ing performance BMIs have predominantly utilized spikes as the decoder's input signals which are recorded using intracortical microelectrode arrays [1]. Despite compelling results reported in several studies [2], [3], spike-based BMIs face two major challenges towards their clinically viable translation [4]. First, the number of recorded spike signals declines over time, which can degrade the performance of BMI decoder. This long-term instability is thought to be caused by scar tissue formation around the electrodes and micromotion of the electrodes [4], [5]. Second, due to high *Nur Ahmadi sampling rate (>10 kHz), detecting and sorting spikes in an implanted circuitry or transmitting the raw data require high power consumption, which will in turn increase heat and limit the device's battery lifetime [4]. One approach to address these challenges is to utilize another signal modality within neural activity, namely local field potential (LFP), as an alternative input signal for BMI decoder. LFP is a low frequency extracellular voltage thought to be mainly generated from postsynaptic currents and reflect activity of population of neurons in the vicinity of the electrodes. LFP is thus believed to be less sensitive to scar formation and micromotion of the electrodes [4]. Several studies have shown that LFPs are more stable than spikes and that considerable amount of movement-related information contained within LFPs are still present even after spikes are lost [1], [6], [7]. In addition, LFPs can be processed at signif- icantly lower sampling rate than spikes, which translates into lower power consumption and lower complexity (i.e. memory requirements). However, despite these interesting properties, LFP-based BMIs have been shown to yield low decoding performances compared to spike-based BMIs [1], [5], [6], [8], [9]. Most of LFP-based BMIs employ Kalman filter (KF) decoder which assumes linear and Gaussian distribution on both the observation and state dynamics. However, since LFPs exhibit nonlinear, non-stationary, and non-Gaussian characteristics [10], aforementioned assumptions could lead to suboptimal results. Therefore, a more flexible and effec- tive decoding method is required to improve the decoding performance of LFP-based BMIs. One promising solution to this issue is recurrent neural networks with long short-term memory (LSTM) architecture, which do not require any assumption on the data. In fact, LSTMs have become the state-of-the-art methods for various applications ranging from speech recognition and synthesis, language modeling and translation, to audio and video anal- ysis [11]. Despite the successes in a variety of fields, to our knowledge, however, LSTMs have not been applied to LFP- based BMI decoding. is supported by the graduate scholarship awarded by Indonesia Endowment Fund for Education (LPDP), Republic of Indonesia. 1Nur Ahmadi, Timothy G. Constandinou, and Christos-Savvas Bouganis are with the Department of Electrical and Electronic Engineering, Imperial College London, SW7 2BT, UK. Email: {n.ahmadi16, t.constandinou, christos-savvas.bouganis}@imperial.ac.uk In light of this, in the present paper, we propose an LSTM- based decoder for decoding hand kinematics from multichan- nel LFPs recorded intracortically from non-human primate (NHP) during free-reaching tasks. We then compare offline decoding performance of the proposed LSTM decoder to a commonly used Kalman filter (KF) decoder. Additionally, we also benchmark the decoding performance between LFP- This preprint is the "accepted" version by IEEE NER. c(cid:13)2019 IEEE. Personal use of this material is permitted. Permission from IEEE must be obtained for all other uses, in any current or future media, including reprinting/republishing this material for advertising or promotional purposes, creating new collective works, for resale or redistribution to servers or lists, or reuse of any copyrighted component of this work in other works. 2Nur Ahmadi and Timothy G. Constandinou are additionally with the Centre for Bio-Inspired Technology, Institute of Biomedical Engineering, Imperial College London, SW7 2AZ, UK Fig. 1. Block diagram of LFP-based BMI decoding using an LSTM network. The larger gray-shaded block illustrates a detailed schematic of an LSTM. (HT). We computed Hilbert envelope only from delta band driven LSTM decoder and KF decoder driven by two types (0-4 Hz) since other frequency bands in our data did not of spike signals: single-unit activity (SUA) and multi-unit contain significant information as observed through time- activity (MUA). frequency decomposition. Both LMP and DHE features were computed within overlapping windows that slide every 4 ms to match the timescale of hand kinematics. We used window widths ranging from 32 ms to 512 ms with an increment of 32 ms. For performance evaluation, we selected a window width and LFP features that yielded the best decoding performance on validation set. The processing steps of our LFP-based BMI decoding is illustrated in Fig. 1. A. Neural Recording and Behavioral Task II. METHODS (dataset Spikes were extracted from each channel by band pass filtering (from 500 to 5000 Hz) raw neural data and then detecting the filtered signal amplitudes that crossed a prede- termined threshold value. All the detected spikes within sin- gle channel are referred to as multi-unit activity (MUA). To extract single-unit activity (SUA), the detected spikes were sorted (i.e. classified) into distinct putative single units. More detailed information on spike detection and sorting processes can be found in [13]. In this study, units with spike rates below 0.5 Hz were excluded [13] thus leaving only 130 (124) SUAs and 91 (94) MUAs for dataset I (II). We computed spike counts within different overlapping bin/window widths (same range as of LFPs) sliding every 4 ms (corresponding to the kinematics time scale). Overlapping bin was used because it has been demonstrated to yield better decoding performance than non-overlap bin [15], [16]. Similar to that of LFP-based decoding, we selected a bin width that resulted in the best decoding performance on validation set. C. Long Short-Term Memory (LSTM) Network Long short-term memory (LSTM) is a type of recurrent neural networks (RNNs) that was developed by Hochreiter and Schmidhuber in 1997 [17]. It has successfully addressed the vanishing or exploding gradient problem encountered when training traditional RNNs. LSTM is very well suited to many sequential data problems as it is capable of learning long-term temporal dependencies through gating mechanism. A commonly used variant of LSTM units contains three gates (forget, input, and output) and a memory cell that control the In this study, we used two datasets I: I 20170124 01; dataset II: I 20170127 03) from a public neural data recorded from primary motor cortex (M1) area of an adult male Rhesus macaque monkey (Macaca mulata) by Sabes Lab [12]. Recording duration for dataset I and II were 9 and 12 minutes, respectively. The neural data was recorded using 96-channel silicon microelectrode array (Utah array with 400 µm inter-electrode distance and 1 mm electrode length) while the monkey was performing self- paced reaching tasks without inter-trial intervals within 8-by- 8 square grid. The recording was referenced to a silver wire placed under the dura (several cm away from the electrodes). The neural data was sampled at 24.4 kHz and filtered with a 4th oder low-pass filter at 7.5 kHz. Fingertip position in x-y coordinates was sampled at 250 Hz. Velocity was then computed from position by using discrete derivative. A more detailed description of the experiment setup is given in [13]. B. Neural Signal Processing LFPs were obtained by low pass filtering raw neural data using 4th order Butterworth filter at 300 Hz and then down- sampling them to 1 kHz. LFPs have been shown to contain common noises, partly arising from the use of a single, distal reference (unipolar) [14]. To remove these common noises, we performed common average reference (CAR) and subtracted it from LFP signal in each channel. At each time instant, CAR was computed by averaging LFP signals across all channels. Two LFP features, local motor potential (LMP) [1], [6], [9] and delta Hilbert envelope (DHE), were extracted. LMPs were computed from LFPs by using time- domain moving average filter. DHEs were calculated firstly by low pass filtering LFPs with 4th order Butterworth filter at 4 Hz and then computing the envelope (i.e. instantaneous amplitude) of their analytic signals through Hilbert transform TABLE I HYPERPARAMETERS FOR LSTM DECODER Hyperparameter Number of units Number of epochs Batch size Dropout rate Learning rate Values {50, 75, 100, · · · , 200} {2, 3, 4, · · · , 8} {32, 64, 96, 128} {0, 0.1, 0.2, 0.3, 0.4} {0.001, 0.0015, 0.002, · · · , 0.003} flow of information. In this study, we used this LSTM variant where its components' state at time instant t can be described by: ft = σ(Wf xt + Uf ht−1 + bf ) it = σ(Wixt + Uiht−1 + bi) ct = tanh(Wcxt + Ucht−1 + bc) ct = ft (cid:12) ct−1 + it (cid:12) ct ot = σ(Woxt + Uoht−1 + bo) ht = tanh(ct) (cid:12) ot (1) where x, h, f, i, o, c represent the input, output, forget gate, input gate, output gate, and memory cell, respectively; σ and tanh denote sigmoid and hyperbolic tangent activation functions, respectively; (cid:12) denotes element-wise multiplica- tion. W, U, b are weight matrices and bias vector parameters that need to be learned during training. Training an LSTM decoder requires setting up configura- tion parameters called hyperparameters, which in this study are given in Table I. The hyperparameter configuration was tuned through a Bayesian optimization library (Hyperopt) [18] with 200 iterations. The LSTM decoder was imple- mented using Keras framework with TensorFlow backend and trained using RMSprop optimizer and root mean squared error (RMSE) loss function. D. Performance Evaluation and Metrics To evaluate the performance of the LSTM decoder, we used two common metrics, namely, root mean squared error (RMSE) and Pearson's correlation coefficient. These metrics were computed between decoded and actual velocities on x- and y- coordinates. We split each dataset into k = 10 subdatasets which were then categorized into a training set (concatenation of k − 2 subdatasets), validation set (1 subdataset), and testing set (1 subdataset). The training set was used to find values of the LSTM parameters (W, U, b). The validation set (from dataset I) was used to find the opti- mal configuration of LSTM hyperparameters, LFP features, window widths, and lags between LFP/spike features and the kinematics that minimized average RMSE across x-y coordinates. This configuration was then used for evaluating the performance of LSTM decoder on testing sets. We iterated the performance evaluation on k different testing sets for each dataset. III. RESULTS According to our evaluation on validation set (dataset I), the following configuration of LSTM hyperparameters led to smallest average RMSE: number of layer = 1, number of timesteps = 2, number of units = 100, number of epochs = 6, Fig. 2. Comparison of decoding performance of LSTM decoder under different LFP features and window widths. Red stars indicate window widths that yielded smallest average RMSEs on validation set of dataset I. batch size = 32, dropout rate = 0.2, and learning rate = 0.001. This configuration was then used to find the optimal LFP features and window width using the validation set (dataset I) and to benchmark the performance of LSTM decoder against KF decoder using the testing sets (dataset I and II). A. LFP Feature and Window Width Selection The LFP features used for comparison were LMP, DHE, and combination of both, which were extracted from two reference schemes (unipolar and CAR). According to our results, CAR-based LFP features led to superior decoding performance than unipolar-based LFP features. LMP, on average, resulted in better performance than DHE. Combi- nation of both LMP and DHE features yielded improved performance than that of single LFP feature. In regards to different window widths ({32, 64, 96,··· , 512} ms), we observed a similar trend across different LFP features. The decoding performance increased (smaller average RMSE) as the window width increased up to a particular point where above this point the performance would decrease. This par- ticular point corresponds to a window width within a range of 256− 320 ms. The summary of LSTM decoder performance with respect to different LFP features and window widths is depicted in Fig. 2. For performance benchmark against KF decoder on testing sets, we selected a combination of LMP and DHE features extracted from CAR-based LFPs with 320 ms window width. Through the same procedure, we selected a combination of LMP and DHE features extracted from unipolar-based LFPs with 256 ms for KF decoder (figure is not shown due to page limitation). B. Window Width Selection for Spike-driven Decoders We also investigated the impact of different window widths to decoding performance of LSTM and KF decoders driven by spike signals (SUA and MUA). A similar trend as in LFP-driven decoders was also observed in the cases of spike-driven LSTM and KF decoders. The decoding performance improved as the window width increased and at a particular point the performance would decline. The 326496128160192224256288320352384416448480512Window Width (ms)37.540.042.545.047.550.052.5Average RMSE (mm/s)UNI[LMP]CAR[LMP]UNI[DHE]CAR[DHE]UNI[LMP+DHE]CAR[LMP+DHE] Fig. 3. Comparison of decoding performance of LSTM decoder driven by SUA and MUA under different window widths. Red stars indicate window widths that yielded smallest average RMSEs on validation set of dataset I. summary of spike-driven LSTM decoder performance with respect to different window widths is shown in Fig. 3. Due to page limitation, the summary of spike-driven KF decoder is not shown. For performance comparison on testing sets, we selected a window width of 256 ms for both SUA- and MUA-driven LSTM decoders. Window widths of 256 ms and 288 ms were selected for SUA- and MUA-driven KF decoders, respectively. C. Decoding Performance Comparison Firstly, we compared the decoding performance of our proposed LFP-driven LSTM decoder to LFP-, SUA-, and MUA-driven KF decoder (referred to as LFP-LSTM, LFP- KF, SUA-KF, and MUA-KF respectively). As shown in Fig. 4, LFP-LSTM achieved significantly better decoding performance than LFP-KF, SUA-KF, and MUA-KF mea- sured in both metrics (RMSE and correlation) for both datasets (I and II). LFP-LSTM yielded average RMSE of 36.90 ± 2.85 (43.60 ± 2.31) mm/s and average correlation of 0.83 ± 0.02 (0.82 ± 0.02) for dataset I (II). This average RMSE value corresponded to performance improvement of 35.59±4.77 (32.93±3.78)% for dataset I (II) with respect to LFP-KF as shown in Fig. 5. In term of average correlation, the performance improvement made by LFP-LSTM over LFP-KF was 17.32 ± 2.85 (14.98 ± 2.65)% for dataset I (II). In addition, we also compared the decoding performance of LFP-LSTM to SUA- and MUA-driven LSTM decoder (referred to as SUA-LSTM and MUA-LSTM). The results show that LFP-LSTM outperformed SUA-LSTM but was outperformed by MUA-LSTM for both datasets. The overall performance comparison among different decoding meth- ods is shown in Fig. 4. We plotted performance improve- ment/decrease of different decoders with respect to LFP- KF as can be seen in Fig. 5. MUA-LSTM yielded largest performance improvement of 40.15±5.07 (36.15±4.42)% in average RMSE and 20.91± 2.40 (17.81± 3.78)% in average correlation for dataset I (II). SUA-KF showed improved performance than LFP-KF in dataset I but worse performance in dataset II. Examples of actual and decoded velocities (in x and y directions) from LFP-LSTM and LFP-KF are given Fig. 4. Decoding performance comparison of LSTM and KF decoders under different input signals. Black horizontal lines inside the boxes repre- sent the medians; red stars represent the means; colored boxes denote the interquartile ranges; whiskers extend 1.5× from upper and lower quartiles. in Fig. 6. For the sake of readability, only decoded velocities from these two decoders are shown. IV. DISCUSSION In this study, we propose a new decoder based on LSTM network to predict (offline) hand kinematics from multichan- nel LFPs of an NHP subject. We have shown that LFP-driven LSTM decoder significantly outperforms KF decoder driven by LFP, SUA, and MUA. This demonstrates the effectiveness of LSTM decoder in capturing nonlinear and complex rela- tionship between LFPs and hand kinematics. On the other hand, KF decoder with its inherent assumptions results in poor performance, especially for LFPs that exhibit high spa- tial correlation [14]. LFPs recorded from unipolar reference are often contaminated by spatially correlated noises, which affects the decoding performance. These noises can be elimi- nated by using common average reference (CAR) employed prior to LFP feature extraction process. Our results show that CAR can lead to improved decoding performance. Our results also show that LMP contains significant amount of movement-related information as indicated by good decoding performance compared to other LFP features, which is in good agreement with [1]. However, our results differs from [1] in that delta band of LFP is more informative than the higher bands. By combining LMP with DHE as the input for LSTM decoder, the performance of LSTM decoder can be improved. Prior studies demonstrated that LFP-driven decoder performed slightly worse or comparable to spike- driven decoder [1], [5]. Using hybrid LFP-spike signals, they could achieve better decoding performance than that of spikes only. Here, we show that using only LFP signals, our proposed LFP-driven LSTM decoder can significantly outperform spike-driven KF decoder. Performance comparison using the same LSTM decoder with different input signals (LFP, SUA, and MUA) show that MUA-LSTM achieves the best decoding performance, 326496128160192224256288320352384416448480512Window Width (ms)3540455055Average RMSE (mm/s)SUA-LSTMMUA-LSTMDataset I30.040.050.060.0Average RMSE (mm/s)LFP-LSTMLFP-KFSUA-LSTMSUA-KFMUA-LSTMMUA-KFDataset II40.050.060.070.0Average RMSE (mm/s)Dataset I0.700.750.800.85Average CorrelationDataset II0.600.700.80Average Correlation Fig. 5. Performance comparison of different decoders relative to LFP-KF. A Positive (negative) value indicates performance improvement (decrease). Black error bars represent 95% confidence intervals. followed by LFP-LSTM and SUA-LSTM, respectively. This may indicate that MUA contains more movement-related information than that of SUA. Better performance of MUA over SUA for both KF and LSTM decoders obtained from this study offers a new perspective to the debate whether or not spikes should be sorted [19], [20]. However, the changes of spike amplitude and waveform over time pose critical challenges in adjusting the threshold crossing value for spike detection, sorting and tracking the spikes into the same putative single units. Furthermore, the high sampling rate required for these spike processing demands high power consumption that hinders the implementation of wireless, scalable, and implantable BMIs. LFP offers attractive proper- ties to address these challenges: long-term stability and low- frequency content (i.e. low sampling rate). In this study, we only use two datasets corresponding two recording sessions (3 days gap between the sessions). For future work, we will investigate the stability and robustness of LFP-, SUA-, and MUA-driven LSTM decoders over long period of time. Overall, our results suggest that LFP indeed contains a rich movement-related information, which corroborates the idea that LFP is a promising alternative signal input for BMIs. Along with their stable and low-frequency properties, LFPs coupled with LSTM decoder could potentially provide high decoding performance, robust, and low power BMIs. ACKNOWLEDGEMENT We thank J. E. O'Doherty and P. N. Sabes for making their data publicly available. REFERENCES [1] S. D. Stavisky et al., "A high performing brain -- machine interface driven by low-frequency local field potentials alone and together with spikes," J. Neural Eng., vol. 12, no. 3, p. 036009, 2015. [2] M. M. Shanechi, "Brain -- machine interface control algorithms," IEEE Trans. Neural Syst. Rehabil. Eng, vol. 25, no. 10, pp. 1725 -- 1734, 2017. [3] D. M. Brandman et al., "Human intracortical recording and neural decoding for brain-computer interfaces," IEEE Trans. Neural Syst. Rehabil. Eng, vol. 25, no. 10, pp. 1687 -- 1696, 2017. Fig. 6. Examples of actual and decoded velocities in x- and y-directions from dataset II. [4] A. Jackson and T. M. Hall, "Decoding local field potentials for neural interfaces," IEEE Trans. Neural Syst. Rehabil. Eng, vol. 25, no. 10, pp. 1705 -- 1714, 2017. [5] E. J. Hwang and R. A. Andersen, "The utility of multichannel local field potentials for brain -- machine interfaces," J. Neural Eng., vol. 10, no. 4, p. 046005, 2013. [6] R. D. Flint et al., "Long term, stable brain machine interface per- formance using local field potentials and multiunit spikes," J. Neural Eng., vol. 10, no. 5, p. 056005, 2013. [7] D. Wang et al., "Long-term decoding stability of local field potentials from silicon arrays in primate motor cortex during a 2D center out task," J. Neural Eng., vol. 11, no. 3, p. 036009, 2014. [8] A. K. Bansal et al., "Decoding 3D reach and grasp from hybrid signals in motor and premotor cortices: spikes, multiunit activity, and local field potentials," J. Neurophysiol., vol. 107, no. 5, pp. 1337 -- 1355, 2011. [9] R. D. Flint et al., "Accurate decoding of reaching movements from field potentials in the absence of spikes," J. Neural Eng., vol. 9, no. 4, p. 046006, 2012. [10] D. L. Menzer et al., "Characterization of trial-to-trial fluctuations in local field potentials recorded in cerebral cortex of awake behaving macaque," J. Neurosci. Methods, vol. 186, no. 2, pp. 250 -- 261, 2010. [11] K. Greff et al., "LSTM: A search space odyssey," IEEE Trans. Neural Syst. Rehabil. Eng, vol. 28, no. 10, pp. 2222 -- 2232, 2017. [12] J. E. O'doherty et al., multichannel sensorimotor doi:10.5281/zenodo.583331. "Nonhuman primate cortex electrophysiology," reaching with 2017, [13] J. G. Makin et al., "Superior arm-movement decoding from cortex with a new, unsupervised-learning algorithm," J. Neural Eng., vol. 15, no. 2, p. 026010, 2018. [14] L. Xinyu et al., "Adaptive common average reference for in vivo multichannel local field potentials," Biomed. Eng. Letters, vol. 7, no. 1, pp. 7 -- 15, 2017. [15] S. Koyama et al., "Comparison of brain -- computer interface decoding algorithms in open-loop and closed-loop control," J. Comput. Neu- rosci., vol. 29, no. 1-2, pp. 73 -- 87, 2010. [16] N. Ahmadi et al., "Spike rate estimation using bayesian adaptive kernel smoother (BAKS) and its application to brain machine interfaces," in Proc. of the Ann. Int. Conf. of the IEEE Engineering in Medicine and Biology Society. IEEE, 2018, pp. 2547 -- 2550. [17] S. Hochreiter and J. Schmidhuber, "Long short-term memory," Neural Computation, vol. 9, no. 8, pp. 1735 -- 1780, 1997. [18] J. Bergstra et al., "Hyperopt: a python library for model selection and hyperparameter optimization," Comput. Sci. Discovery, vol. 8, no. 1, p. 014008, 2015. [19] S. Todorova et al., "To sort or not to sort: the impact of spike-sorting on neural decoding performance," J. Neural Eng., vol. 11, no. 5, p. 056005, 2014. [20] B. P. Christie et al., "Comparison of spike sorting and thresholding of voltage waveforms for intracortical brain -- machine interface perfor- mance," J. Neural Eng., vol. 12, no. 1, p. 016009, 2014. Dataset I−10010203040Δ Average RMSE (%)LFP-KFLFP-LSTMSUA-LSTMSUA-KFMUA-LSTMMUA-KFDataset II−10010203040Δ Average RMSE (%)Dataset I−1001020Δ Correlation (%)Dataset II−1001020Δ Correlation (%)0246810−2000200x-velocity (mm/s)ActualLFP-KFLFP-LSTM0246810Time (s)−400−2000200y-velocity (mm/s)
1907.11570
1
1907
2019-07-26T13:31:26
Decoding the circuitry of consciousness: from local microcircuits to brain-scale networks
[ "q-bio.NC" ]
Identifying the physiological processes underlying the emergence and maintenance of consciousness is one of the most fundamental problems of neuroscience, with implications ranging from fundamental neuroscience to the treatment of patients with disorders of consciousness (DOC). One major challenge is to understand how cortical circuits at drastically different spatial scales, from local networks to brain-scale networks, operate in concert to enable consciousness, and how those processes are impaired in DOC patients. In this review, we attempt to relate available neurophysiological and clinical data with existing theoretical models of consciousness, while linking the micro- and macro-circuit levels. First, we address the relationships between awareness and wakefulness on the one hand, and cortico-cortical, and thalamo-cortical connectivity on the other hand. Second, we discuss the role of three main types of GABAergic interneurons in specific circuits responsible for the dynamical re-organization of functional networks. Third, we explore advances in the functional role of nested oscillations for neural synchronization and communication, emphasizing the importance of the balance between local (high-frequency) and distant (low-frequency) activity for efficient information processing. The clinical implications of these theoretical considerations are presented. We propose that such cellular-scale mechanisms could extend current theories of consciousness.
q-bio.NC
q-bio
Decoding the circuitry of consciousness: from local microcircuits to brain-scale networks Julien Modolo1, Mahmoud Hassan1, Fabrice Wendling1,*, and Pascal Benquet1 1: Univ Rennes, INSERM, LTSI -- U1099, F-35000 Rennes, France *Corresponding author: Fabrice Wendling -- [email protected] Abstract Identifying the physiological processes underlying the emergence and maintenance of consciousness is one of the most fundamental problems of neuroscience, with implications ranging from fundamental neuroscience to the treatment of patients with disorders of consciousness (DOC). One major challenge is to understand how cortical circuits at drastically different spatial scales, from local networks to brain-scale networks, operate in concert to enable consciousness, and how those processes are impaired in DOC patients. In this review, we attempt to relate available neurophysiological and clinical data with existing theoretical models of consciousness, while linking the micro- and macro-circuit levels. First, we address the relationships between awareness and wakefulness on the one hand, and cortico-cortical, and thalamo-cortical connectivity on the other hand. Second, we discuss the role of three main types of GABAergic interneurons in specific circuits responsible for the dynamical re-organization of functional networks. Third, we explore advances in the functional role of nested oscillations for neural synchronization and communication, emphasizing the importance of the balance between local (high-frequency) and distant (low-frequency) activity for efficient information processing. 1 The clinical implications of these theoretical considerations are presented. We propose that such cellular-scale mechanisms could extend current theories of consciousness. Keywords Disorders of consciousness, functional connectivity, micro-circuitry, communication through coherence, gating by inhibition, electroencephalography. Introduction Understanding how consciousness arises from communication among brain regions is a question of the utmost importance in the field of neuroscience in general, and for the diagnosis and treatment of patients suffering from disorders of consciousness (DOC) in particular. The problem of consciousness can be seen as fundamental (e.g.: "What is consciousness? Why do we have subjective, conscious experiences?", such questions are referred to as the "hard" problem of consciousness (Harnad, 1998)) or more empirical (e.g.: "What are the processes associated with the emergence and maintenance of consciousness?", this forms the "soft" problem of consciousness (Harnad, 1998)). In this review, we aim at understanding 1) how brain networks at different scales are involved in enabling and maintaining conscious processes of information transmission and processing related to awareness and wakefulness, and 2) how these mechanisms are related to the disruptions of consciousness in DOC patients. Many theories have been proposed to explain how consciousness originates, ranging from abstract and informational concepts to neurophysiology-based theories. The most widespread theories of consciousness have a fundamental assumption in common: information processing in 2 the human brain networks is inextricably linked with consciousness. A recent paper by Dehaene and colleagues summarizes this principle as follows (Dehaene et al., 2017): "What we call "consciousness" results from specific types of information-processing computations, physically realized by the hardware of the brain." The three main theories of consciousness include the Integrated Information Theory (IIT) (Tononi, 2004), the Dynamic Core Hypothesis (Tononi and Edelman, 1998) (DCH), and the Global Workspace Theory (Baars, 1988; Dehaene et al., 1998; Dehaene et al., 2003) (GWT). Historically, DCH theory has been the first to refer to the notion of information processing involved in consciousness (Tononi and Edelman, 1998). This theory is based on the central role of functional clusters in the thalamo-cortical system and re-entrant interactions, with high integration and differentiation of neuronal activity being crucial in the emergence of conscious phenomena. IIT, which is an evolution and generalization of DCH, is based on a set of axioms from which postulates are derived. IIT also provides a computable quantity, F, also called integrated information, that quantifies the level of consciousness. In this framework, if combining sub-elements increases information processing capability more than linearly adding these elements, then integrated information increases. Global Workspace Theory (GWT) is a theory of consciousness theory that is more directly connected with neurophysiology and neuroanatomy. The main hypothesis of GWT is that conscious information is globally available within the brain, and that two fundamentally different computational systems co-exist: 1) a network of distributed "local" processors operating in parallel in the brain ("unconscious"), and 2) a "global" workspace formed by a network of distributed interconnected cortical areas involved in conscious perception (Baars, 1988). The key concept here is that the global workspace is composed of distant regions densely connected through glutamatergic cortico- 3 cortical connections as opposed to the network of local processors operating in "isolation" (in parallel). It is worth noting that this distinction between unconscious and conscious processes has been recently challenged, and might be an oversimplification (Melnikoff and Bargh, 2018). In GWT, conscious perception is associated with "ignition", a large-scale brain activation pattern induced by exposure to a stimulus (Dehaene et al., 2003). If the stimulus does not trigger ignition, and if the induced brain response remains spatially confined and is brief, then the perception will not reach consciousness. In other terms, a stimulus has to be sufficiently long and strong to reach consciousness, which suggests a form of filtering mechanism that is consistent with the view that only a limited amount of information effectively enters in the global workspace. Despite these successes in accounting for experimental data in humans regarding subliminal (unconscious) and conscious perception (Sergent and Dehaene, 2004; King et al., 2016), one drawback of GWT is that it does not explicitly relate the large-scale recruitment of brain regions during conscious access with cellular mechanisms. More precisely, what prevents ignition for short, irrelevant stimuli; and conversely, what enables ignition for strong stimuli? The neuroanatomical, neurophysiological and dynamical mechanisms behind ignition are of fundamental importance to understand how we become conscious of a stimulus, or how alterations of brain networks can result in impaired consciousness in DOC patients. If one accepts that consciousness is associated with a sufficiently complex (in the algorithmic sense of "less compressible") information processing, then the emergence of consciousness is critically dependent on three factors: 1) a physical network enabling interactions between its components; 2) the flexibility to re-organize transiently sub-networks to achieve greater computation capabilities by increasing the number of possible configurations and input-output functions, through functional connectivity; and 3) dynamic communications between its 4 components. These three critical components have the potential to be altered, for example in lesions following traumatic brain injury. While the physical large-scale network linking brain regions is well defined and known as the connectome, there are still unresolved questions regarding the transient organization of clusters performing specific computations (functional networks), the associated means of communication (neural coding) and how large-scale functional brain networks and information routing can reconfigure rapidly depending on micro- circuits regulation. In this review, the objective is therefore to propose a bridging between cortical microcircuits on the one hand (cellular scale), and brain-scale activity associated with the two main dimensions of conscious perception (awareness and wakefulness) on the other hand. Such multi- scale understanding is a pre-requisite to understand how brain networks become dysfunctional in DOC, and might contribute to reconcile GWT and DCH into a unified framework. The review is organized as follows. First, we examine the relationship between the two dimensions of consciousness, namely wakefulness and awareness, with functional connectivity between cortical regions and the thalamus. Second, we review the "means of communication" enabling complex information processing linked with consciousness, which regulate cortico- cortical communication, among which communication through coherence (CTC) and gating by inhibition (GBI). The alteration of those mechanisms is presented through results from the clinical literature. Third, we attempt at linking these findings with concepts that have recently emerged based on the communication between brain regions based on cross-frequency couplings between oscillations with specific functional roles. Finally, we suggest possible clinical implications of this framework in terms of novel neuromodulation protocols in DOC. 5 1. Awareness, wakefulness: a short review of concepts Conscious perception results from an interplay between two processes interacting with each other: awareness and wakefulness. Deep sleep switches awareness off, whereas being able of conscious perception (awareness) of environmental stimuli usually implies a state of wakefulness, as illustrated in Figure 1. For example, during general anaesthesia, there is both an absence of conscious perception, awareness and wakefulness. In some peculiar cases, however, these two components can be unrelated. Unresponsive Wakefulness State (UWS) is an example of Disorder of Consciousness (DOC) in which wakefulness is present without any detectable signs of awareness (Laureys and Boly, 2012). Also, during lucid dreaming, there is a form of awareness in the absence of wakefulness (during sleep) (Voss et al., 2013). Another example is spatial neglect syndrome, in which patients have no conscious awareness of visual stimuli, while being awake in the contralateral side of the cortical lesion (Le et al., 2015). Awareness is supported by attentional, fronto-parietal networks that amplify synaptic connections within specific cortical pathways (Tallon-Baudry, 2011). This amplification of relevant stimuli enhances the activated network related to stimulus representation. In parallel, the concomitant inhibition of irrelevant surrounding networks i) optimizes cortico-cortical routing of information by constraining the possible propagation of neural activity throughout all possible cortical "routes", which ii) restricts propagation to a limited number of stimulus-driven possibilities, and iii) increases the signal-to-noise ratio. Such mechanisms are related to the concept of functional connectivity, and are detailed further in this review. Importantly, such mechanisms of active inhibition likely involve cortical inhibition with an active modulation by thalamocortical inputs (Gabernet et al., 2005), implying that the pattern of thalamo-cortical activity influences information processing in cortico-cortical networks. 6 Wakefulness depends critically on thalamo-cortical connectivity and neuromodulatory brainstem inputs to the thalamus (including noradrenaline projections from the locus coeruleus (Monti, 2011)). For instance, during slow-wave sleep, a low-frequency, synchronized activity between the cortex and thalamus (the so-called "up-and-down" rhythm (Neske, 2015), prevents transmission of sub-cortical inputs to the cortex during sleep). This provides an example in which thalamo-cortical drastically decreases information processing by cortico-cortical networks, resulting in a loss of consciousness. During wakefulness, thalamo-cortical activity is weakly synchronized (Gent et al., 2018), which is a necessary, but not sufficient condition to enable consciousness. For example, as aforementioned, wakefulness is present in UWS patients but cortico-cortical communication is severely impaired (Noirhomme et al., 2010), interfering with the "awareness" component of consciousness. Another required condition for consciousness is an efficient large-scale cortico-cortical communication that can support awareness through the activation of attentional fronto-parietal networks (Luckmann et al., 2014; Ptak et al., 2017). Therefore, in terms of neuroanatomy, it is possible to link wakefulness with thalamo-cortical, "vertical" connectivity, while awareness depends on cortico-cortical, "horizontal" connectivity. This is consistent with the recent view by Naccache (Naccache, 2018) that Minimally Conscious State (MCS) patients, who are conscious to some degree, exhibit "Cortically Mediated States", whereas UWS patients do not exhibit such activity, presumably because cortico-cortical connectivity (critical for awareness) is too severely impaired. More specifically, a "critical mass" of information processing requires occurring for the emergence and maintenance of consciousness, which is tightly regulated by thalamo-cortical and cortico-cortical connectivity. Figure 1 presents, in a two-dimensional plane, the continuum of the states of consciousness, as a function of awareness and wakefulness. 7 Figure 1. Wakefulness and awareness are two essential dimensions of consciousness. In this diagram, several qualitatively different states of consciousness have been positioned on the 2D-matrix as a function of the associated axes "content of consciousness" (awareness) and "level of consciousness" (wakefulness). Adapted from (Laureys, 2005). 2. Functional networks, a flexible architecture for conscious processes The main disadvantage of static network architectures is their limitation in terms of amount and variety (complexity) of information processing that can take place. The brain takes advantage of different mechanisms that overcome this limitation, by enabling a dynamic, transient reconfiguration of brain networks increasing the repertoire of possible responses to 8 inputs (i.e., complexity of input-output functions) (Sporns, 2013). Such transient networks involving only a few brain regions, coordinated to achieve a specific function limited in duration, form what is termed functional connectivity. There is a growing interest regarding the functional networks associated with specific cognitive tasks (Hassan et al., 2015) and novel frameworks have recently emerged (Avena-Koenigsberger et al., 2017) to explain how brain-scale anatomical connectivity relates to functional connectivity. Functional networks organize through the network "means of communication", also termed communication dynamics, that governs information routing through specific networks, instead of propagating information through the entire brain network (Avena-Koenigsberger et al., 2017). If that was the case, then information generated locally would induce distant activity in all connectome-related regions, resulting in a massively synchronized response with low informational content and complexity. More specifically, one fundamental question is: what are the mechanisms regulating communication dynamics and enabling functional networks to emerge in brain-scale networks? This question is central to understand how the brain optimizes its information processing capabilities, which are tightly linked with consciousness. We propose that the fundamental mechanisms underlying communication dynamics are actually cellular-scale mechanisms that 1) prevent brain-scale neuronal synchronization following a stimulus, and 2) enable the transient coupling of specific distant brain regions. There has been a considerable amount of interest for large-scale brain activity patterns linked with consciousness, since those can be measured through various neuroimaging modalities (e.g., electroencephalography, EEG; functional magnetic resonance imaging, fMRI). However, mechanisms at the cellular scale have remained more challenging to address in humans for obvious reasons of invasiveness associated with the 9 required recording techniques in humans. In this section, we review evidence for such mechanisms that could bridge the micro-circuit and brain-scale levels. In terms of large-scale neuroanatomical pathways enabling consciousness, long-range glutamatergic projections between pyramidal neurons through white matter fibers have likely a critical role (Dehaene and Changeux, 2011) since they enable fast (due to myelin) communication between distant regions. At the brain-scale level, these white matter fibers are likely critical to enable conscious access, which involves the transient stabilization of neuronal activity encoding a specific information pattern, in a network of high-level brain regions interconnected by long-range connections, with the prefrontal cortex (PFC) acting as a key node (Dehaene et al., 2006; Berkovitch et al., 2017). On conscious trials, distributed gamma-band activity reflects a stabilization of local information broadcasted to other areas. Global broadcasting is thought to make the information accessible to introspection and reportable to other brain regions (Lamme, 2010). During conscious access to a specific information, other surrounding global workspace neurons would be inhibited and unavailable for the processing of other stimuli, therefore remaining preconscious (not reaching consciousness). At the local scale, a micro-circuit has also been identified as being involved in the communication between distant brain regions: the projection from pyramidal neurons in a brain region to VIP-positive (Vasoactive Intestinal Peptide) neurons in another region. By activating VIP-positive neurons in a distant region, this induces an inhibition of somatostatin-positive (SST) neurons, which target pyramidal cells dendrites, resulting in a disynaptic disinhibition (Karnani et al., 2016). Through this disynaptic disinhibition, gamma activity generation can occur through PV-PV mutual inhibition, and binding between the two involved regions can 10 possibly take place (Munoz et al., 2017), temporarily enabling information transfer and processing. These cellular-scale mechanisms are summarized in Figure 2. Figure 2. Basic (schematic) circuits involved in the generation of local and distant oscillations. In a local cortical network (cortical column), gap-junctional, mutual inhibition of soma-projecting, "fast" GABAergic interneurons are one of the basic mechanisms of generation for local gamma activity, along with the PING (Pyramid-InterNeuron Gamma). Conversely, the feedback loop between dendrite-projecting, "slow" GABAergic interneurons and pyramidal cells can generate low-frequency activity. Importantly, distant communication through the disynaptic pathway enabling transient generation of gamma oscillations in distant populations. Pyramidal cells in the source population (left circuit) project on the pyramidal cells of the distant population (right), but also on VIP interneurons that project on dendrite-projecting SST neurons. Transient activation of VIP neurons from the source population transiently inhibits SST neurons in the target population, enabling the generation of gamma oscillations through the PV-PV and PYR-PV circuit. Once the input on distant VIP neurons decreases, SST neurons resume their inhibitory input, which can terminate gamma oscillations generation. 11 Considering these cellular-scale mechanisms, one crucial question is to understand how they are involved into the transient emergence and maintenance of functional networks such as those supporting consciousness. It is established that the state of consciousness is critically dependent on brain functional cortico-cortical connectivity (Marino et al., 2016; Naro et al., 2018). As a reminder, functional connectivity refers to statistically significant couplings between temporal courses of neuronal activity within different regions while anatomical connectivity denotes the physical connections between brain regions. Functional networks can therefore reflect indirect connections between brain regions, and are transient depending on which tasks are performed, or which stimuli are perceived. Even in the absence of any specific task or stimuli, it has been shown that resting-state networks (e.g., default mode network -DMN-) are also transient (Kabbara et al., 2017). Experimental evidence supports the idea that functional connectivity can shed light on the networks involved in various conscious states (Jin and Chung, 2012). For example, during general anaesthesia-induced loss of consciousness, there is a breakdown in cortico-cortical functional connectivity (Ferrarelli et al., 2010; Hudetz, 2012; Gomez et al., 2013), severely impairing the capacity of cortical networks to integrate information and to make it available at a large scale, as required for conscious perception in IIT or GWT. Similarly, in the transition from wakefulness to slow-wave sleep, the firing rate in the cortex remains relatively unchanged during the depolarizing phases of the slow sleep oscillation (Steriade et al., 2001), while effective brain connectivity is dramatically altered (Tononi and Sporns, 2003; Esser et al., 2009). Upon falling into NREM sleep, cortical activations also become more local and stereotypical, impairing effective cortical connectivity (Massimini et al., 2010), as shown using TMS-evoked EEG responses which remain very close to the stimulation site; while these responses involve a network of distant brain regions undergoing complex dynamical patterns of 12 successive activation during wakefulness (Guillery and Sherman, 2002; Casali et al., 2013; Casarotto et al., 2016). These results also emphasize the crucial role of the thalamo-cortical pathway in cortico-cortical functional connectivity. It is worth noting that the essential role of the thalamocortical loop as well as so-called "reentrant interactions" were previously considered as key in the DCH (Tononi and Edelman, 1998). Consistently with these results obtained during sleep, this breakdown of cortico-cortical connectivity has also been observed during general anaesthesia and in DOC patients, and even explored through computational modeling (Esser et al., 2009; Casali et al., 2013). In the GWT framework, this explains why consciousness is impaired in such states: large-scale communication between distant brain areas is impaired due to thalamo-cortical modulation, preventing ignition from occurring. In brain-damaged DOC patients, large-scale cortico-cortical communication can be impaired through the partial destruction of long-range fibers, physically impeding long-range brain synchrony. In terms of effects at the cellular scale, destruction of long-range fibers could prevent the synchronization of distant VIP interneurons, which is critical to induce disynaptic disinhibition and associated gamma activity required for CTC. Pathological alterations of functional connectivity have been investigated using a variety of modalities: (1) functional connectivity during ''resting state'' using fMRI or EEG; (2) pulsed stimulation using transcranial magnetic stimulation (TMS) during EEG recording; and (3) other perturbation-based approaches investigating brain responses to sensory stimuli (Boly et al., 2017). The advantage of functional connectivity is that it can be employed to improve the evaluation and classification of disorders of consciousness (Sanders et al., 2012; Holler et al., 2014; Rossi Sebastiano et al., 2015; Naro et al., 2018). For example, in mild cognitive impairment (MCI) patients, it has been shown that impaired consciousness is associated with 13 altered effective connectivity (Varotto et al., 2014; Crone et al., 2015). Failure of large-scale connectivity, along with a hypersynchrony of local short-range delta and alpha activity were detected within the DMN and were correlated with the level of awareness in patients with DOC (Vanhaudenhuyse et al., 2010; Fingelkurts et al., 2013; Maki-Marttunen et al., 2013; Varotto et al., 2014; Kabbara et al., 2017; Naro et al., 2018). Furthermore, the functional connectivity pattern of several brain regions, such as the posterior cingulate cortex and precuneus, may even predict UWS patients' state improvement of consciousness with an accuracy superior to 80% (Wu et al., 2015). Beside "passive" investigation of resting-state functional connectivity, the use of TMS- evoked EEG responses enables the active "probing" of functional connectivity. For example, a drastic breakdown of functional connectivity has been identified in UWS patients using a specific TMS protocol triggering, in these patients, a stereotyped, local EEG response similar as in unconscious sleeping or anaesthetized subjects (Rosanova et al., 2012). Restoring cortical large-scale effective connectivity with transcranial brain stimulation, such as tACS, in DOCs could therefore be a useful approach to facilitate partial recovery by enhancing oscillations and plasticity. One clinical result supporting this idea is the recent demonstration that DLPFC (dorsolateral prefrontal cortex)-tACS was able to transiently restore the connectivity breakdown in DOC individuals (Naro et al., 2016). One fundamental microscopic-scale mechanism involved in information routing in the brain, and contributing to form functional networks within the anatomical network, is called Gating By Inhibition (Jensen and Mazaheri, 2010). GBI involves inhibitory processes resulting in the selective activation of sub-networks and inactivation of other sub-networks. By preventing brain- scale activation in response to a stimulus, and restricting the number of brain regions engaged in 14 performing tasks, GBI also prevents states of low complexity (e.g., all brain regions displaying the exact same activity) and therefore inefficient information processing. Thus, GBI processes suggest that the role of inhibition is more complex than preventing excessive activation of brain networks, contributing instead to shaping anatomical brain networks into functional networks (Avena-Koenigsberger et al., 2017). Possible alterations of GBI were reported in studies showing that EEG alpha power is lower in UWS than in MCS patients (Lehembre et al., 2012; Stefan et al., 2018), hinting that the neurobiological mechanisms underlying alpha oscillations generation and associated GBI are profoundly altered in unresponsive patients. Moreover, alpha activity was highly synchronized and clustered in central and posterior cortical regions in UWS patients (Lehembre et al., 2012; Stefan et al., 2018), suggesting a possible failure of GBI in the most severe disorders of consciousness. 3. Information processing in large-scale functional networks through nested oscillations One of the most established processes by which distant brain regions engage together in an activity pattern associated with the performance of a given task is Communication Through Coherence (CTC) (Fries, 2005, 2009, 2015b; Deco and Kringelbach, 2016; Bonnefond et al., 2017). CTC involves indeed phase-coupled gamma activity between distant brain regions to enable information processing. CTC has been suggested to be the substrate of "binding", i.e. the merging of different features of a stimulus into a single, unified conscious perception (Singer, 2001). More precisely, the excitability fluctuation in a group of neurons provides a specific signature characterized by a specific frequency band and pattern of discharge (Womelsdorf et al., 2014), propagating through a large-scale network consisting of anatomically interconnected 15 brain areas and subsequently triggering activity in connected regions. Information processing in the brain is strongly linked with phase-locked, coordinated-in-time fluctuations of excitability (Fries, 2005; Fries, 2015a) in networks of distributed neuronal populations. The resulting oscillations generate a specific neuronal code, and coherence enables the association of information and communication. Furthermore, CTC involves gamma activity, generated mainly by GABAergic interneurons (PV-positive basket cells). Taken together, inhibitory processes appear key for information routing and processing in brain-scale networks involved in consciousness: GBI shapes brain networks spatially (which brain regions are involved, and which ones are inhibited), while CTC controls them temporally (information flow). However, this raises an intriguing question: if gamma rhythms are generated locally by interneuronal - GABAergic- networks, how can distant brain regions, connected through glutamatergic long- range fibers, communicate efficiently and achieve CTC? One possibility is that the co- occurrence of low- and high-frequency neuronal oscillations could provide distant co-activation (low-frequency, glutamatergic origin) that would then enable binding (high-frequency, GABAergic origin). This would involve the formation of a functional network of several brain regions through the low-frequency rhythm, prior to information transfer and processing through CTC (involving gamma activity). As a support for this possibility, conscious perception is indeed characterized by an increase in distributed gamma-band activity (Melloni et al., 2007; Wyart and Tallon-Baudry, 2009). Interestingly, these fast oscillations are modulated by slow oscillations (Osipova et al., 2008; Jensen et al., 2014). It has recently been proposed that phase synchronization of low-frequency oscillations, playing the role of a temporal reference frame for information, carrying high- 16 frequency activity, is a general mechanism of brain communication (Bonnefond et al., 2017). These nested oscillations might be a key mechanism, not only for cortico-cortical communication and processing, but also between sub-cortical structures. Emotional memory, involving both cortical and subcortical structures, indeed engages large network synchronization through nested theta-gamma oscillations (Bocchio et al., 2017). During in vivo experiments performed in rodents, a perceived threat (a stimulus announcing a footshock) enhances theta power and coherence in the amygdala, prefrontal cortex, and hippocampus (Lesting et al., 2011; Likhtik et al., 2014), while fast gamma bursts are phase-locked to theta oscillations (Stujenske et al., 2014). Overall, these results support the idea that nested oscillations at theta and gamma frequencies are a plausible substrate for information channel opening/routing (theta) and processing/transfer (gamma) within the brain. Attention is another key element for conscious processing and it is involved in the synchronization of distant brain regions (Steinmetz et al., 2000; Niebur et al., 2002). The main underlying brain rhythms involved in attentional processes are alpha and gamma oscillations: brain regions synchronize gamma oscillations (Womelsdorf et al., 2014), and are modulated by slow alpha oscillations. Slow oscillations enable inhibition of irrelevant networks, influence local signal processing, widespread information exchange, and perception (Sadaghiani and Kleinschmidt, 2016). Information flow is established by neuronal synchronization at the lower- frequency bands, namely in the theta (4 -- 7 Hz), alpha (8 -- 13 Hz), and beta (14 -- 25 Hz) bands (Bonnefond et al., 2017). One possible reason is that low-frequency activity induces a transient change in excitability in target brain structures, which provides an optimal window for binding neuronal signals from different regions through high-frequency activity (i.e., gamma) (Canolty et al., 2006). This provides further support to the idea that low-frequency neural oscillations are 17 mainly involved in establishing transient long-range communication through glutamatergic projections, while high-frequency neural oscillations are rather involved in information processing/transfer. It is therefore possible to relate the notion of "integration" with this long- range, glutamatergic co-activation, enabling brain-scale communication between brain regions; while "differentiation/segregation" would rather depend on locally generated gamma activity (and in part on low-frequency activation level, which would result in massively synchronized activity, and reduced differentiation and complexity). An overview of the aforementioned mechanisms is proposed in Figure 3. Figure 3. Schematic overview of transient, selective binding among cortical networks through cellular-scale mechanisms. A. Schematic diagram of an anatomical network with main projections between regions at the brain scale. B. Upper panel. Selective binding in subset of cortical regions occurs through the generation of gamma 18 oscillations (mostly through micro-circuits involving basket cells), while the distant disinhibition of specific brain regions occurs through disynaptic disinhibition (activation of distant VIP neurons inhibits SST neurons, which in turn decrease their inhibitory projection on pyramidal neurons). This contributes to shape the anatomical network into a functional network. The alpha rhythm acts as a pulsed inhibition to inhibit "irrelevant" networks, increasing further the signal-to-noise (SNR) ratio. Lower panel. Decreased integration (e.g., following brain damage) leads to impaired synchronization of distant brain regions (reflected by decreased low-frequency rhythm on the illustrative oscillation), and thereby a decrease of binding, which combined to a decrease in GBI efficiency strongly decreases the overall SNR, leading to dysfunction of the network in terms of integration and binding required for consciousness. Let us note that, in addition to decreased amplitude of the low-frequency rhythm, the phase relationship of the nested gamma oscillation could be perturbed (i.e., more random) as compared to the physiological case. Such potential relationship of nested theta/gamma oscillations remains to be explored in DOC. Emerging evidence shows that the local versus global information processing balance can be impaired in neurological disorders. Typically, a recent study investigating functional networks in Alzheimer's disease patients identified a decrease in brain integration as quantified by the participation coefficient (reflects communication between distant brain modules), while segregation as quantified by the clustering coefficient (reflects local communication between neighbour brain regions) was increased (Kabbara et al., 2018). This is consistent with neurodegenerative processes, which likely impact the "locking" of specific brain regions or the inhibition of irrelevant networks, thereby severely impairing large-scale integration of information. In the context of DOC, recent clinical evidence (Chennu et al., 2017; Rizkallah et al., 2019) supports this view. In the study by Chennu et al., scalp-level networks were assessed from DOC patients and pointed at decreased integration within the alpha band. More specifically, the fronto-parietal network in the alpha band was discriminant between MCS and UWS patients. In a recent study, Rizkallah et al. (Rizkallah et al., 2019) quantified the level of local versus global information processing in frequency-dependent functional networks (source 19 level) in DOC patients and controls. Integration in theta band functional networks decreased with consciousness level, and two anatomical regions were systematically involved between controls and any patient group: a portion of the left orbitofrontal cortex and the left precuneus. One possibility is that physical damage to long-range white matter fibers impairs large-scale integration and the local/global information processing balance. One possible approach to study such anatomical damage to white matter fibers is diffusion tensor imaging (DTI), as performed in DOC patients suffering severe brain injury (Fernandez-Espejo et al., 2011; Galanaud et al., 2012; Luyt et al., 2012), which highlighted widespread disruptions of white matter. Lower fractional anisotropy was indeed found in the subcortico-cortical and cortico-cortical fiber tracts of DOC patients as compared to controls (Lant et al., 2016; Weng et al., 2017), suggesting that major consciousness deficits in DOC patients may be related to altered WM connections between the basal ganglia, thalamus, and frontal cortex. This is also in line with the effect of lesion of myelinated fiber tracts, which can result in a failure of communication between distant brain regions (Adams et al., 2000). Therefore, it seems reasonable that white matter lesions can alter, modify or prevent both CTC and GBI between large-scale networks. Furthermore, we speculate that, should the specific phase-locking of gamma oscillations onto theta oscillations be perturbed, then clinical manifestations associated with DOC might appear (loss of integration and decrease in the consciousness level). 4. Possible clinical implications In this review, we have attempted to reconcile the neuroanatomical and neurophysiological knowledge at the level of micro- and macro-scopic networks, regarding the processes that underlie the emergence and maintenance of consciousness, and its alterations in DOC patients. 20 The multiplexing of neuronal rhythms through nested oscillations appears as a plausible mechanism of co-activation in a network of specific distant brain regions (integration), which is a pre-requisite for a conscious perception. Furthermore, a key mechanism seems to be the subtle balance between low-frequency activity (associated with "global" processing) and high- frequency activity (associated with a more "local" processing), which could enable the neuronal dynamics underlying optimal information routing and processing. Excessive low-frequency activity (e.g., delta activity) results in massively, synchronized activity resulting in a loss of complexity in terms of information processing, paralleled in such cases with a loss of consciousness (e.g., sleep, seizures). Similarly, a lack of fronto-parietal functional coupling (attentional network) has been recently observed in a recent study, as quantified using high- resolution EEG in DOC patients (Chennu et al., 2017), suggesting that a sufficient level of fronto-parietal coupling is required to achieve sufficient neuronal integration and ignition for conscious perception. More generally, brain dynamics in DOC patients is typically characterized by a loss of integration at a large-scale (Chennu et al., 2017; Rizkallah et al., 2019), preventing efficient large-scale coordination of distant brain regions to achieve conscious perception. This suggests that the low-frequency rhythm required for long-range cortical communication is decreased, preventing binding in the gamma range and therefore further processing information and ignition for consciousness access. That being said, what are the possible implications of this slow/fast activity balance as a candidate mechanism for the complex processing associated with consciousness? An interesting perspective, which would also be a form of validation for this mechanism, is the use of neuromodulation techniques in DOC patients to increase their level of consciousness. The objective of such neuromodulation techniques could be to "re-balance" local versus global 21 processing, for example through the use of transcranial direct or alternating stimulation (tDCS/tACS), applied to both a frontal and a parietal site simultaneously, in order to increase low-frequency synchronization in the theta range. It is the current view that tACS can modulate endogenous brain rhythms using relatively low-levels electric fields (<1 V/m, as discussed in (Modolo et al., 2018), which would involve to use a stimulation frequency in the theta range to increase residual oscillations in this frequency range (i.e., assuming that residual anatomical connections are still present). Dual-site, fronto-parietal tACS in the theta range could then provide a non-invasive possibility to increase the level of consciousness in DOC patients, pending that some residual anatomical connectivity remains in the case of brain-damaged patients. Interestingly, a recent study (Violante et al., 2017) used dual-site tACS in the theta range in healthy volunteers reported improved working memory, a function also dependent on fronto-parietal networks. Another study, using tACS targeting the fronto-temporal network, reported an increase in working memory performance in seniors to comparable levels than young participants (Reinhart and Nguyen, 2019). These recent results, obtained in humans, provide compelling evidence that re-balancing information processing through neuromodulation protocols could contribute to increase the level of consciousness in some DOC patients (e.g., where damage to long-range white matter fibers is not too severe). Discussion and concluding remarks The two dimensions of consciousness, awareness and wakefulness, depend on anatomically distinct pathways: cortico-cortical "horizontal" connectivity and thalamo-cortical "vertical" connectivity. Excessive "up-and-down-like" thalamo-cortical activity impairs cortico-cortical connectivity by due to excessive lateral inhibition, thereby preventing CTC of distant brain 22 regions, resulting in drastically altered functional connectivity, in line with the loss of consciousness in deep sleep or DOC. This control of cortico-cortical communication by thalamo- cortical activity is fundamental in understanding how attentional processes can emerge by transiently recruiting efficiently, through nested low-(theta) and high-(gamma) frequency rhythms, distant brain regions. The evidence reviewed highlights how the balance of nested brain rhythms with fundamentally different functions can transform an anatomical network into a transient, successive activation of different sub-networks, i.e. a functional network. This versatility of reconfiguration of the structural connectome results in an immense and complex dynamical repertoire of functional networks with specific rhythms and cross-frequency couplings, probably a key infrastructure enabling consciousness. Among those rhythms, three appear especially involved in conscious processes: while the function of the theta rhythm appears linked with "opening" transient channels of communications through distant regions, the alpha rhythm seems to play the role of pulsed inhibition to increase further the SNR. There is also solid and converging evidence that gamma oscillations are an excellent candidate for information processing and transfer. Importantly, mechanisms identified at the micro-circuit scale between specific types of interneurons, such as projections from pyramidal cells to distant VIP cells, are critical to provide a more mechanistic framework for theories of consciousness, notably GWT. Such mechanisms indeed clarify the conditions under which ignition can occur, while providing links with other concepts that are not necessarily unified (e.g., CTC, GBI) to enable access to consciousness. In addition to the recruitment of selective brain regions to have access to the global workspace, it also appears important to take into account that active inhibitory processes co-occur to improve the signal-to-noise ratio (e.g., GBI). An overview of those concepts, along with the contribution 23 of the reviewed mechanisms to the increase in neural activity complexity associated with consciousness, is provided in Figure 4. Figure 4. Synthesis of the network-level mechanisms underlying the complexity associated with conscious processes. The structural characteristics of brain circuits at different scales are mentioned with some key oscillatory rhythms and associated functions. Taken together, the mechanisms presented in this review suggest that the importance of the thalamo-cortical pathway, emphasized in the DCH theory, cannot be neglected within the context of the GW theory: the thalamo-cortical pathway plays actually the role of a "switch", enabling or 24 not efficient integration and communication within cortico-cortical networks through feed- forward inhibition. Therefore, efficient modulation of the thalamo-cortical pathway is necessary, but not sufficient, to enable ignition within cortical networks and availability of information at a large scale. For these reasons, it seems that DCH and GWT are both accurate each from their perspective, and could be unified to obtain a more integrated vision, through a new framework accounting for both aspects (ignition, availability of information, and control/routing of information by thalamo-cortical pathways). In such framework, two balances are critical: the first one is between vertical (thalamo-cortical) and horizontal (cortico-cortical) connectivity, which controls the second one between local and distant information processing within cortico-cortical networks. Consequently, we propose that DCH and GWT could be reconciled through this balance between horizontal and vertical connectivity, and account for a wider range of phenomena related to consciousness and its deregulations. Future directions - An important step forward would be to investigate further the cross-frequency coupling between the low-frequency theta rhythm and high-frequency gamma rhythm in healthy controls as compared to DOC patients during resting state. The identification of such changes could have implications in terms of diagnostic evaluation, but also regarding novel neuromodulation protocols that might aim, at least in part, to regulate abnormal cross-frequency couplings. - Another promising application would be to translate the circuitry presented in this review into a tractable computational model consisting in a network of brain regions, possibly using the 25 neural mass model approach. Evaluating in silico how the microcircuits are involved into the generation of nested theta-gamma oscillations, and how TMS-evoked EEG responses at the brain scale are impacted by synchronized thalamocortical activity, would provide key mechanistic understanding. Candidate tDCS/tACS protocols could also be tested and evaluated in silico. - Characterizing further the dynamics of functional networks in DOC patients using EEG, for example by studying the nature and dynamics of modular states over time (e.g., dwell time) as a function of the level of consciousness (wake, sleep, DOC such as UWS). Extracting such dynamical information about functional brain network could have diagnostic implications, notably to distinguish between MCS and UWS patients. 26 Competing financial interests statement The authors have no competing financial interests to declare for this work. Funding This study is funded by the Future Emerging Technologies (H2020-FETOPEN-2014-2015- RIA under agreement No. 686764) as part of the European Union's Horizon 2020 research and training program 2014 -- 2018. References Adams JH, Graham DI, Jennett B. The neuropathology of the vegetative state after an acute brain insult. Brain 2000; 123 ( Pt 7): 1327-38. Avena-Koenigsberger A, Misic B, Sporns O. Communication dynamics in complex brain networks. Nat Rev Neurosci 2017; 19(1): 17-33. Baars BJ. A cognitive theory of consciousness. Cambridge England ; New York: Cambridge University Press; 1988. Berkovitch L, Dehaene S, Gaillard R. Disruption of Conscious Access in Schizophrenia. Trends Cogn Sci 2017; 21(11): 878-92. Bocchio M, Nabavi S, Capogna M. Synaptic Plasticity, Engrams, and Network Oscillations in Amygdala Circuits for Storage and Retrieval of Emotional Memories. Neuron 2017; 94(4): 731-43. Boly M, Massimini M, Tsuchiya N, Postle BR, Koch C, Tononi G. Are the Neural Correlates of Consciousness in the Front or in the Back of the Cerebral Cortex? Clinical and Neuroimaging Evidence. The Journal of neuroscience : the official journal of the Society for Neuroscience 2017; 37(40): 9603-13. Bonnefond M, Kastner S, Jensen O. Communication between Brain Areas Based on Nested Oscillations. eNeuro 2017; 4(2). Canolty RT, Edwards E, Dalal SS, Soltani M, Nagarajan SS, Kirsch HE, et al. High gamma power is phase-locked to theta oscillations in human neocortex. Science 2006; 313(5793): 1626-8. Casali AG, Gosseries O, Rosanova M, Boly M, Sarasso S, Casali KR, et al. A theoretically based index of consciousness independent of sensory processing and behavior. Sci Transl Med 2013; 5(198): 198ra05. Casarotto S, Comanducci A, Rosanova M, Sarasso S, Fecchio M, Napolitani M, et al. Stratification of unresponsive patients by an independently validated index of brain complexity. Annals of neurology 2016; 80(5): 718-29. Chennu S, Annen J, Wannez S, Thibaut A, Chatelle C, Cassol H, et al. Brain networks predict metabolism, diagnosis and prognosis at the bedside in disorders of consciousness. Brain 2017; 140(8): 2120-32. Crone JS, Schurz M, Holler Y, Bergmann J, Monti M, Schmid E, et al. Impaired consciousness is linked to changes in effective connectivity of the posterior cingulate cortex within the default mode network. Neuroimage 2015; 110: 101-9. Deco G, Kringelbach M. Metastability and Coherence: Extending the Communication through Coherence Hypothesis Using a Whole-Brain Computational Perspective. Trends Neurosci 2016; 39(6): 432. Dehaene S, Changeux JP. Experimental and theoretical approaches to conscious processing. Neuron 2011; 70(2): 200-27. 27 Dehaene S, Changeux JP, Naccache L, Sackur J, Sergent C. Conscious, preconscious, and subliminal processing: a testable taxonomy. Trends Cogn Sci 2006; 10(5): 204-11. Dehaene S, Kerszberg M, Changeux JP. A neuronal model of a global workspace in effortful cognitive tasks. Proc Natl Acad Sci U S A 1998; 95(24): 14529-34. Dehaene S, Lau H, Kouider S. What is consciousness, and could machines have it? Science 2017; 358(6362): 486- 92. Dehaene S, Sergent C, Changeux JP. A neuronal network model linking subjective reports and objective physiological data during conscious perception. Proc Natl Acad Sci U S A 2003; 100(14): 8520-5. Esser SK, Hill S, Tononi G. Breakdown of effective connectivity during slow wave sleep: investigating the mechanism underlying a cortical gate using large-scale modeling. Journal of neurophysiology 2009; 102(4): 2096- 111. Fernandez-Espejo D, Bekinschtein T, Monti MM, Pickard JD, Junque C, Coleman MR, et al. Diffusion weighted imaging distinguishes the vegetative state from the minimally conscious state. Neuroimage 2011; 54(1): 103-12. Ferrarelli F, Massimini M, Sarasso S, Casali A, Riedner BA, Angelini G, et al. Breakdown in cortical effective connectivity during midazolam-induced loss of consciousness. Proc Natl Acad Sci U S A 2010; 107(6): 2681-6. Fingelkurts AA, Fingelkurts AA, Bagnato S, Boccagni C, Galardi G. Dissociation of vegetative and minimally conscious patients based on brain operational architectonics: factor of etiology. Clin EEG Neurosci 2013; 44(3): 209-20. Fries P. A mechanism for cognitive dynamics: neuronal communication through neuronal coherence. Trends Cogn Sci 2005; 9(10): 474-80. Fries P. Neuronal gamma-band synchronization as a fundamental process in cortical computation. Annu Rev Neurosci 2009; 32: 209-24. Fries P. Rhythms for Cognition: Communication through Coherence. Neuron 2015a; 88(1): 220-35. Fries P. Rhythms for Cognition: Communication through Coherence. Neuron 2015b; 88(1): 220-35. Gabernet L, Jadhav SP, Feldman DE, Carandini M, Scanziani M. Somatosensory integration controlled by dynamic thalamocortical feed-forward inhibition. Neuron 2005; 48(2): 315-27. Galanaud D, Perlbarg V, Gupta R, Stevens RD, Sanchez P, Tollard E, et al. Assessment of white matter injury and outcome in severe brain trauma: a prospective multicenter cohort. Anesthesiology 2012; 117(6): 1300-10. Gent TC, Bandarabadi M, Herrera CG, Adamantidis AR. Thalamic dual control of sleep and wakefulness. Nat Neurosci 2018. Gomez F, Phillips C, Soddu A, Boly M, Boveroux P, Vanhaudenhuyse A, et al. Changes in effective connectivity by propofol sedation. PLoS One 2013; 8(8): e71370. Guillery RW, Sherman SM. Thalamic relay functions and their role in corticocortical communication: generalizations from the visual system. Neuron 2002; 33(2): 163-75. Harnad S. Explaining consciousness: the hard problem. Trends Cogn Sci 1998; 2(6): 234-5. Hassan M, Benquet P, Biraben A, Berrou C, Dufor O, Wendling F. Dynamic reorganization of functional brain networks during picture naming. Cortex 2015; 73: 276-88. Holler Y, Thomschewski A, Bergmann J, Kronbichler M, Crone JS, Schmid EV, et al. Connectivity biomarkers can differentiate patients with different levels of consciousness. Clin Neurophysiol 2014; 125(8): 1545-55. Hudetz AG. General anesthesia and human brain connectivity. Brain Connect 2012; 2(6): 291-302. Jensen O, Gips B, Bergmann TO, Bonnefond M. Temporal coding organized by coupled alpha and gamma oscillations prioritize visual processing. Trends Neurosci 2014; 37(7): 357-69. Jensen O, Mazaheri A. Shaping functional architecture by oscillatory alpha activity: gating by inhibition. Front Hum Neurosci 2010; 4: 186. Jin SH, Chung CK. Messages from the brain connectivity regarding neural correlates of consciousness. Exp Neurobiol 2012; 21(3): 113-22. Kabbara A, Eid H, El Falou W, Khalil M, Wendling F, Hassan M. Reduced integration and improved segregation of functional brain networks in Alzheimer's disease. J Neural Eng 2018; 15(2): 026023. Kabbara A, El Falou W, Khalil M, Wendling F, Hassan M. The dynamic functional core network of the human brain at rest. Sci Rep 2017; 7(1): 2936. Karnani MM, Jackson J, Ayzenshtat I, Hamzehei Sichani A, Manoocheri K, Kim S, et al. Opening Holes in the Blanket of Inhibition: Localized Lateral Disinhibition by VIP Interneurons. The Journal of neuroscience : the official journal of the Society for Neuroscience 2016; 36(12): 3471-80. King JR, Pescetelli N, Dehaene S. Brain Mechanisms Underlying the Brief Maintenance of Seen and Unseen Sensory Information. Neuron 2016; 92(5): 1122-34. Lamme VA. How neuroscience will change our view on consciousness. Cogn Neurosci 2010; 1(3): 204-20. 28 Lant ND, Gonzalez-Lara LE, Owen AM, Fernandez-Espejo D. Relationship between the anterior forebrain mesocircuit and the default mode network in the structural bases of disorders of consciousness. Neuroimage Clin 2016; 10: 27-35. Laureys S. The neural correlate of (un)awareness: lessons from the vegetative state. Trends Cogn Sci 2005; 9(12): 556-9. Laureys S, Boly M. Unresponsive wakefulness syndrome. Arch Ital Biol 2012; 150(2-3): 31-5. Le A, Stojanoski BB, Khan S, Keough M, Niemeier M. A toggle switch of visual awareness? Cortex 2015; 64: 169- 78. Lehembre R, Marie-Aurelie B, Vanhaudenhuyse A, Chatelle C, Cologan V, Leclercq Y, et al. Resting-state EEG study of comatose patients: a connectivity and frequency analysis to find differences between vegetative and minimally conscious states. Funct Neurol 2012; 27(1): 41-7. Lesting J, Narayanan RT, Kluge C, Sangha S, Seidenbecher T, Pape HC. Patterns of coupled theta activity in amygdala-hippocampal-prefrontal cortical circuits during fear extinction. PLoS One 2011; 6(6): e21714. Likhtik E, Stujenske JM, Topiwala MA, Harris AZ, Gordon JA. Prefrontal entrainment of amygdala activity signals safety in learned fear and innate anxiety. Nat Neurosci 2014; 17(1): 106-13. Luckmann HC, Jacobs HI, Sack AT. The cross-functional role of frontoparietal regions in cognition: internal attention as the overarching mechanism. Prog Neurobiol 2014; 116: 66-86. Luyt CE, Galanaud D, Perlbarg V, Vanhaudenhuyse A, Stevens RD, Gupta R, et al. Diffusion tensor imaging to predict long-term outcome after cardiac arrest: a bicentric pilot study. Anesthesiology 2012; 117(6): 1311-21. Maki-Marttunen V, Diez I, Cortes JM, Chialvo DR, Villarreal M. Disruption of transfer entropy and inter- hemispheric brain functional connectivity in patients with disorder of consciousness. Front Neuroinform 2013; 7: 24. Marino S, Bonanno L, Giorgio A. Functional connectivity in disorders of consciousness: methodological aspects and clinical relevance. Brain Imaging Behav 2016; 10(2): 604-8. Massimini M, Ferrarelli F, Murphy M, Huber R, Riedner B, Casarotto S, et al. Cortical reactivity and effective connectivity during REM sleep in humans. Cogn Neurosci 2010; 1(3): 176-83. Melloni L, Molina C, Pena M, Torres D, Singer W, Rodriguez E. Synchronization of neural activity across cortical areas correlates with conscious perception. The Journal of neuroscience : the official journal of the Society for Neuroscience 2007; 27(11): 2858-65. Melnikoff DE, Bargh JA. The Mythical Number Two. Trends Cogn Sci 2018. Modolo J, Denoyer Y, Wendling F, Benquet P. Physiological effects of low-magnitude electric fields on brain activity: advances from in vitro, in vivo and in silico models. Current Opinion in Biomedical Engineering 2018; 8: 38-44. Modolo J, Thomas AW, Legros A. Neural mass modeling of power-line magnetic fields effects on brain activity. Front Comput Neurosci 2013; 7: 34. Monti JM. Serotonin control of sleep-wake behavior. Sleep Med Rev 2011; 15(4): 269-81. Munoz W, Tremblay R, Levenstein D, Rudy B. Layer-specific modulation of neocortical dendritic inhibition during active wakefulness. Science 2017; 355(6328): 954-9. Naccache L. Minimally conscious state or cortically mediated state? Brain 2018; 141(4): 949-60. Naro A, Bramanti A, Leo A, Cacciola A, Manuli A, Bramanti P, et al. Shedding new light on disorders of consciousness diagnosis: The dynamic functional connectivity. Cortex 2018; 103: 316-28. Naro A, Bramanti P, Leo A, Russo M, Calabro RS. Transcranial Alternating Current Stimulation in Patients with Chronic Disorder of Consciousness: A Possible Way to Cut the Diagnostic Gordian Knot? Brain Topogr 2016; 29(4): 623-44. Neske GT. The Slow Oscillation in Cortical and Thalamic Networks: Mechanisms and Functions. Front Neural Circuits 2015; 9: 88. Niebur E, Hsiao SS, Johnson KO. Synchrony: a neuronal mechanism for attentional selection? Curr Opin Neurobiol 2002; 12(2): 190-4. Noirhomme Q, Soddu A, Lehembre R, Vanhaudenhuyse A, Boveroux P, Boly M, et al. Brain connectivity in pathological and pharmacological coma. Front Syst Neurosci 2010; 4: 160. Osipova D, Hermes D, Jensen O. Gamma power is phase-locked to posterior alpha activity. PLoS One 2008; 3(12): e3990. Ptak R, Schnider A, Fellrath J. The Dorsal Frontoparietal Network: A Core System for Emulated Action. Trends Cogn Sci 2017; 21(8): 589-99. Reinhart RMG, Nguyen JA. Working memory revived in older adults by synchronizing rhythmic brain circuits. Nat Neurosci 2019; 22(5): 820-7. 29 Rizkallah J, Annen J, Modolo J, Gosseries O, Benquet P, Mortaheb S, et al. Decreased integration of EEG source- space networks in disorders of consciousness. Neuroimage Clinical 2019; 23(101841): 1-9. Rosanova M, Gosseries O, Casarotto S, Boly M, Casali AG, Bruno MA, et al. Recovery of cortical effective connectivity and recovery of consciousness in vegetative patients. Brain 2012; 135(Pt 4): 1308-20. Rossi Sebastiano D, Panzica F, Visani E, Rotondi F, Scaioli V, Leonardi M, et al. Significance of multiple neurophysiological measures in patients with chronic disorders of consciousness. Clin Neurophysiol 2015; 126(3): 558-64. Sadaghiani S, Kleinschmidt A. Brain Networks and alpha-Oscillations: Structural and Functional Foundations of Cognitive Control. Trends Cogn Sci 2016; 20(11): 805-17. Sanders RD, Tononi G, Laureys S, Sleigh JW. Unresponsiveness not equal unconsciousness. Anesthesiology 2012; 116(4): 946-59. Sergent C, Dehaene S. Is consciousness a gradual phenomenon? Evidence for an all-or-none bifurcation during the attentional blink. Psychol Sci 2004; 15(11): 720-8. Singer W. Consciousness and the binding problem. Ann N Y Acad Sci 2001; 929: 123-46. Sporns O. Structure and function of complex brain networks. Dialogues Clin Neurosci 2013; 15(3): 247-62. Stefan S, Schorr B, Lopez-Rolon A, Kolassa IT, Shock JP, Rosenfelder M, et al. Consciousness Indexing and Outcome Prediction with Resting-State EEG in Severe Disorders of Consciousness. Brain Topogr 2018. Steinmetz PN, Roy A, Fitzgerald PJ, Hsiao SS, Johnson KO, Niebur E. Attention modulates synchronized neuronal firing in primate somatosensory cortex. Nature 2000; 404(6774): 187-90. Steriade M, Timofeev I, Grenier F. Natural waking and sleep states: a view from inside neocortical neurons. Journal of neurophysiology 2001; 85(5): 1969-85. Stujenske JM, Likhtik E, Topiwala MA, Gordon JA. Fear and safety engage competing patterns of theta-gamma coupling in the basolateral amygdala. Neuron 2014; 83(4): 919-33. Tallon-Baudry C. On the neural mechanisms subserving consciousness and attention. Front Psychol 2011; 2: 397. Tononi G. An information integration theory of consciousness. BMC Neurosci 2004; 5: 42. Tononi G, Edelman GM. Consciousness and complexity. Science 1998; 282(5395): 1846-51. Tononi G, Sporns O. Measuring information integration. BMC Neurosci 2003; 4: 31. Vanhaudenhuyse A, Noirhomme Q, Tshibanda LJ, Bruno MA, Boveroux P, Schnakers C, et al. Default network connectivity reflects the level of consciousness in non-communicative brain-damaged patients. Brain 2010; 133(Pt 1): 161-71. Varotto G, Fazio P, Rossi Sebastiano D, Duran D, D'Incerti L, Parati E, et al. Altered resting state effective connectivity in long-standing vegetative state patients: an EEG study. Clin Neurophysiol 2014; 125(1): 63-8. Violante IR, Li LM, Carmichael DW, Lorenz R, Leech R, Hampshire A, et al. Externally induced frontoparietal synchronization modulates network dynamics and enhances working memory performance. Elife 2017; 6. Voss U, Schermelleh-Engel K, Windt J, Frenzel C, Hobson A. Measuring consciousness in dreams: the lucidity and consciousness in dreams scale. Conscious Cogn 2013; 22(1): 8-21. Wendling F, Bartolomei F, Bellanger JJ, Chauvel P. Epileptic fast activity can be explained by a model of impaired GABAergic dendritic inhibition. The European journal of neuroscience 2002; 15(9): 1499-508. Weng L, Xie Q, Zhao L, Zhang R, Ma Q, Wang J, et al. Abnormal structural connectivity between the basal ganglia, thalamus, and frontal cortex in patients with disorders of consciousness. Cortex 2017; 90: 71-87. Womelsdorf T, Valiante TA, Sahin NT, Miller KJ, Tiesinga P. Dynamic circuit motifs underlying rhythmic gain control, gating and integration. Nat Neurosci 2014; 17(8): 1031-9. Wu X, Zou Q, Hu J, Tang W, Mao Y, Gao L, et al. Intrinsic Functional Connectivity Patterns Predict Consciousness Level and Recovery Outcome in Acquired Brain Injury. The Journal of neuroscience : the official journal of the Society for Neuroscience 2015; 35(37): 12932-46. Wyart V, Tallon-Baudry C. How ongoing fluctuations in human visual cortex predict perceptual awareness: baseline shift versus decision bias. The Journal of neuroscience : the official journal of the Society for Neuroscience 2009; 29(27): 8715-25. 30
1805.08239
2
1805
2018-11-26T15:36:21
The Roles of Supervised Machine Learning in Systems Neuroscience
[ "q-bio.NC", "cs.LG", "stat.ML" ]
Over the last several years, the use of machine learning (ML) in neuroscience has been rapidly increasing. Here, we review ML's contributions, both realized and potential, across several areas of systems neuroscience. We describe four primary roles of ML within neuroscience: 1) creating solutions to engineering problems, 2) identifying predictive variables, 3) setting benchmarks for simple models of the brain, and 4) serving itself as a model for the brain. The breadth and ease of its applicability suggests that machine learning should be in the toolbox of most systems neuroscientists.
q-bio.NC
q-bio
The Roles of Supervised Machine Learning in Systems Neuroscience Joshua I. Glaser​1​*, Ari S. Benjamin​1​*, Roozbeh Farhoodi​1​*, Konrad P. Kording​1,2 1. Department of Bioengineering, University of Pennsylvania 2. Department of Neuroscience, University of Pennsylvania, Member, Canadian Institute For Advanced Research *These authors contributed equally to this work. Contacts: jglaser8, ​aarrii, roozbeh, kording @seas.upenn.edu Abstract: Over the last several years, the use of machine learning (ML) in neuroscience has been rapidly increasing. Here, we review ML's contributions, both realized and potential, across several areas of systems neuroscience. We describe four primary roles of ML within neuroscience: 1) creating solutions to engineering problems, 2) identifying predictive variables, 3) setting benchmarks for simple models of the brain, and 4) serving itself as a model for the brain. The breadth and ease of its applicability suggests that machine learning should be in the toolbox of most systems neuroscientists. Introduction: There is a lot of enthusiasm about machine learning (ML). After all, it has allowed computers to surpass human-level performance at image classification ​(He et al. 2015)​, to beat humans in complex games such as "Go" ​(Silver et al. 2016)​, and to provide high-quality speech to text (Hannun et al. 2014) in popular mobile phones. Progress in ML is also getting attention in the scientific community. Writing in the July 2017 issue of Science focusing on "AI Transforms Science", editor Tim Appenzeller writes, "For scientists, prospects are mostly bright: AI promises to supercharge the process of discovery" ​(Appenzeller 2017)​. The field of systems neuroscience is no exception. In the last few years there have been many opinion pieces about the importance of ML in neuroscience ​(Vu et al. 2018; Barak 2017; Paninski and Cunningham 2017; Vogt 2018; Hinton 2011)​. Moreover, when we analyze the number of journal articles about ML in neuroscience, we find that its use has been continuously growing over the last 20 years (Fig. 1). Machine learning has been used in many different ways within this literature. In this review, we will catalog the conceptual applications of ML in systems neuroscience. Figure 1: Growth of Machine Learning in Neuroscience. Here we plot the proportion of neuroscience papers that have used ML over the last two decades. That is, we calculate the number of papers involving both neuroscience and machine learning, normalized by the total number of neuroscience papers. Neuroscience papers were identified using a search for "neuroscience" on Semantic Scholar. Papers involving neuroscience and machine learning were identified with a search for "machine learning" and "neuroscience" on Semantic Scholar. On the highest level, ML is typically divided into the subtypes of supervised, unsupervised, and reinforcement learning. Supervised learning builds a model that predicts outputs from input data. Unsupervised learning is concerned with finding structure in data, e.g. clustering, dimensionality reduction, and compression. Reinforcement learning allows a system to learn the best actions based on the reward that occurs at an end of a sequence of actions. This review focuses on supervised learning. Why is creating progressively more accurate regression or classification methods (see Box 1) worthy of a title like 'The AI Revolution' ​(Appenzeller 2017)​? It is because countless questions can be framed in this manner. When classifying images, an input picture can be used to predict the object in the picture. When playing a game, the setup of the board (input) can be used to predict an optimal move (output). When texting on our smartphones, our current text is used to create suggestions of the next word. Similarly, science has many instances where we desire to make predictions from measured data. In this review, we categorize the ways in which supervised ML promises to assist, or has already been applied to, problems in systems neuroscience. We believe that applications of supervised ML in this field can be divided in roughly four categories (Fig. 2). 1) ​Solving engineering problems​. Machine learning can improve the predictive performance of methods used by neuroscientists, such as medical diagnostics, brain-computer interfaces, and research tools. 2) ​Identifying predictive variables​. Machine learning can more accurately determine whether variables (e.g., those related to the brain and outside world) predict each other. 3) Benchmarking simple models​. We can compare the performance of simple interpretable models to highly accurate ML models in order to help determine the quality of the simple models. 4) Serving as a model for the brain. We can argue whether the brain solves problems in a similar way to ML systems, e.g. deep neural networks. The logic behind each of these applications is rather distinct. For the bulk of this review, we will go into further detail about these four roles of ML in neuroscience. We will provide many examples, both realized and potential, of ML across several areas of systems neuroscience. More specifically, we will discuss the four roles of ML in relation to ​neural function​, including neural activity and how it relates to behavior; and ​neural structure​, i.e., neuroanatomy. We also discuss ML in practice (Box 2), as that is crucial for useful applications. Figure 2: Examples of the four roles of supervised machine learning in neuroscience. 1 - ​ML can solve engineering problems​. For example, it can help researchers control a prosthetic limb using brain activity. 2 - ​ML can identify predictive variables​. For example, by using MRI data, we can identify which brain regions are most predictive for diagnosing Alzheimer's disease ​(Lebedev et al. 2014)​. 3 - ​ML can benchmark simple models​. For example, we can compare the predictive performance of the simple "population vector" model of how neural activity relates to movement (Georgopoulos, Schwartz, and Kettner 1986) to a ML benchmark (e.g. an RNN). 4 - ​ML can serve as a the brain​. For example, researchers have studied how neurons in the visual pathway model of correspond to units in an artificial network that is trained to classify images ​(Yamins and DiCarlo 2016)​. Role 1: Solving engineering problems A surprisingly wide array of engineering problems can be cast as prediction problems. Their common thread is that one would like to estimate some quantity of interest (​Y​​) and can take measurements (​X​​) that relate to that quantity. The relationship between ​X and ​Y​​, however, is unknown and might be complicated. We call these 'engineering problems' when the final quantity, ​Y​​, is all that is desired. In these problems, one does not need detailed understanding of the relationship - the aim is simply to estimate ​Y as accurately as possible. For example, email providers wish to filter spam from incoming messages, and only care about how accurately the emails are sorted. Traditionally, one would attempt to carefully understand the relationship between ​X and ​Y​​, and synthesize this into a model. Modern machine learning (ML) is changing this paradigm. Instead of detailed expert knowledge of a process, a practitioner simply needs a large database of measurements along with the associated quantity of interest for each measurement. Machine learning algorithms can then automatically model their relationship. Once trained, an ML algorithm can make predictions for new measurements. This 'engineering' framework is the traditional application of ML and is common in industry. In neuroscience, there are many problems that resemble this general problem format. Neural Activity / Function Many medical applications depend on successfully extracting information about intention, sensation, or disease from measurements of neural activity. This is a difficult problem, since the meaning of neural activity, the 'neural code', is most often not known. Machine learning is now a ubiquitous tool for this task in situations in which it is possible to obtain large datasets of neural activity along with the behaviors or diseases of interest. One such application is brain-computer interfaces (BCIs), which seek to use neural signals to control prosthetic limbs, computer cursors, or other external objects. Several groups have used modern ML techniques, such as recurrent neural networks, to improve BCIs using recordings of spikes ​(Sussillo et al. 2016, 2012)​, ECoG (Elango et al. 2017)​, or EEG ​(Bashivan et al. 2015)​. Machine learning can also be used to predict future neural activity from past neural activity. This application is relevant for epilepsy, since imminent seizures can be predicted using deep learning ​(Talathi 2017; Nigam and Graupe 2004) and ensemble methods ​(Brinkmann et al. 2016)​. line of applications, researchers have used ML to diagnose neurological conditions from activity. Several reviews on this specific application have recently been published (see ​(Arbabshirani et al. 2017) for classification using neuroimaging data, ​(Rathore et al. 2017) for classification of Alzheimer's disease, and ​(Vieira, Pinaya, and Mechelli 2017) for a focus on deep learning approaches). Due to the large datasets available for neural recordings, ML has improved the standards of accuracy on these medical applications despite the complexity of the input signals. In another An ML approach also promises to assist with the inverse of the above problem: predicting neural activity from variables in the outside world. Solving this problem is important if we want to use neural stimulation to induce accurate sensation. A prosthetic eye, for example, could be built by stimulating retinal ganglion cells in the correct manner according to the output of a camera ​(Nirenberg and Pandarinath 2012)​. The most accurate model of ganglion cell activity is currently a deep learning model trained to predict activity from naturalistic scenes ​(McIntosh et al. 2016)​. Similarly, prosthetic limbs could provide tactile and proprioceptive sensations if somatosensory neurons were correctly stimulated ​(Armenta Salas et al. 2018)​. Machine learning models may help to enable these neural prosthetics to induce sensations. A very similar approach can be used to quantify behavior ​(D. J. Anderson and Perona 2014)​, such as movement, sleeping, and socializing. For example, we may want to quantify movement of the whole body using cheap video recordings. Recent progress in the field has made video quantification far more precise. Researchers have used deep learning to estimate human poses from video ​(Insafutdinov et al. 2016)​. Related approaches have recently gotten easier to use and less data intensive, and have been extended to animal tracking ​(Mathis et al. 2018; T. Pereira et al. 2018)​. Along with estimating poses, we can also directly estimate types of behavior (e.g., walking vs. stopping vs. jumping) from video ​(Kabra et al. 2013)​. Behavior can also be estimated based on other modalities such as audio recordings ​(Arthur et al. 2013)​. This progress in ML-based behavior tracking is timely, given recent calls to understand neural control of more naturalistic behavior in more natural environments ​(Krakauer et al. 2017)​. The engineering approach of applying ML is also helping solve the problem of obtaining accurate estimates of neural activity from raw measurements. Many imaging methods, such as EEG, MEG, and fMRI, require the solving of an 'inverse problem' -- obtaining the source from the measurements. For example, researchers estimate the origin of EEG signals within the brain based on electrode recordings from the scalp. Recently, it has been observed that deep learning can improve the estimates for imaging ​(McCann, Jin, and Unser 2017)​. Neural networks have improved image denoising ​(Burger, Schuler, and Harmeling 2012; Xie, Xu, and Chen 2012) and deconvolution ​(L. Xu et al. 2014)​, can provide super-resolution images ​(Dong et al. 2016)​, and can even replace the entire image processing pipeline ​(Golkov et al. 2016; Vito et al. 2005)​. Outside of imaging, the deconvolution of time series data is another common application. For example, once a researcher has obtained traces of cellular calcium concentration, there is still the difficult 'inverse problem' of inferring the timing of the underlying spiking. Competition-style ML on labeled datasets ​(Theis et al. 2016) provides good solutions (Berens et al. 2017)​. In each of these applications, a difficult engineering problem was replaced by building a large labeled dataset and using ML to learn the desired relationship. Neuroanatomy / Structure Just as neural activity can be indicative of disease, so can neuroanatomy. As such, it is often possible to take anatomical measurements and use machine learning to diagnose disease. For example, researchers can distinguish between Alzheimer's disease and healthy brains of older adults using MRI scans ​(Sarraf, Tofighi, and Others 2016)​. More generally, neuroanatomical measurements such as structural MRI and diffusion tensor imaging (DTI) can distinguish healthy from unhealthy patients across many conditions including schizophrenia, depression, autism and ADHD ​(Arbabshirani et al. 2017; Rathore et al. 2017; Vieira, Pinaya, and Mechelli 2017)​. Sometimes, ML enables surprising opportunities. For example, using deep convolutional neural networks, we can surprisingly predict cardiovascular risk factors from retinal fundus photographs (Poplin et al. 2018)​. The future will undoubtedly see ongoing efforts to automatically detect disease from biological data. Since the majority of research in neuroanatomy is based on imaging techniques, recent advances in computer vision using ML are becoming important tools for neuroanatomy. Thus, segmenting and labeling parts of an image, which usually requires manual annotation, is an especially important task. However, as imaging techniques improve and the volume of data increases, it will become infeasible to rely on manual annotation. To solve this problem, many ML techniques have been developed to automatically segment or label new images based on a techniques in this developing dataset of previously labeled images. The vast majority of literature are based, at least in part, on convolutional neural networks ​(Litjens et al. 2017; Ronneberger, Fischer, and Brox 2015)​. This approach has been employed to label medical images, such as in identifying white matter tracts from MRI scans ​(Ghafoorian et al. 2017)​. They have also been used to understand the connections and morphologies of neurons from electron microscopy ​(Helmstaedter et al. 2013; Funke et al. 2017; Turaga et al. 2010)​. As imaging data improve in resolution and volume, ML is becoming a crucial and even necessary tool for reconstructing and mapping neuroanatomy. Caveats While ML methods have been able to provide many engineering solutions, ML is not magic (Wolpert and Macready 1997)​. Several conditions must be met for an ML method to solve a problem successfully. A first consideration is that the selected method must match the structure of the data. For example, convolutional neural networks make the assumption that images have common local features (like edges), which allows them to be more successful than standard feedforward neural networks. In practice, this means that a user of ML must take care to select their method, or alternatively to preprocess their data, such that the assumptions match how the inputs relate to outputs. This requires good knowledge of the specifics of the data and also of ML methods. A workaround to this is to apply automated ML methods, which iterate intelligently through many possible model configurations and select the best-performing option ​(Guyon et al. 2015; Elsken, Metzen, and Hutter 2018; Feurer et al. 2015; Kotthoff et al. 2017)​. This approach is slow but often works comparatively well. The general rule, however, is that good ML engineering requires a good awareness of the data. is not Another potential failure mode of ML is "overfitting" to the training data. Ideally, an ML method trained on. If a method learns to make accurate should accurately predict data it predictions on the training data but cannot generalize to new data, it is said to be overfit. To guard against the worry of overfitting, ML practitioners always report a model's performance on a test set that the model has not been trained on. The capacity of a method to overfit to data can be lowered with regularization methods, which penalize the complexity of a model. Still, overfitting is especially worrisome for small datasets and complex models. While different methods have different sensitivities to the number of data points, all methods become less vulnerable to overfitting when datasets are large. Sometimes simpler methods may be better choices on small datasets, even if a more complex method could better express the underlying input/output relationship (if there was sufficient data). The risk of overfitting means that all ML practitioners must be aware of regularization techniques, their dataset size, and the importance of reporting accuracy on a test set not used for training. Yet another practical drawback of ML is that it can be slow. For large datasets and complex models, the time it takes to train the model can be prohibitive without proper hardware, like GPUs for deep learning. Once a model is much faster to make predictions. Still, for applications that require real-time predictions, even the prediction step may be too slow for some ML methods. For example, predictions for brain machine interfaces often need to be made in the timescale of tens of milliseconds, which can be a challenge for models requiring many computations. This tradeoff between complexity and run-time is an important aspect in choosing a model for many engineers. is trained, however, it Role 2: Identifying predictive variables Neuroscientists often ask questions of the form, "which variables are related to something of interest?" For example, which brain regions can predict each other? Which brain regions contain information related to a subject's decision? Which cell types are affected by a certain disease? Machine learning (ML) can help to more accurately identify how informative one set of variables is about another. This is particularly instructive when there is a complex nonlinear relationship between the variables, which is often the case in neural systems. Answering these types of questions allows researchers to better understand the relationship between parts of the brain, stimuli, behavior, and more. The general strategy resembles that of the engineering applications (Role 1). However, instead of only searching for maximal predictive accuracy, one examines which input variables lead to that accuracy. There are many methods to establish so-called 'feature importance' (also known as 'feature selection'). Two of the simplest are the leave-one-out strategy, in which each variable is removed and one observes the decrease in accuracy, and the 'best first' strategy, in which the algorithm is run on each variable alone. Leave-one-out reflects the information in that variable but not the others, while best-first reflects the total (learnable) task information in each variable. Both are related but different definitions of what it means for a feature to be important. The development of feature importance measures is an active field in ML and statistics ​(Tang, Alelyani, and Liu 2014)​. These methods allow us to get insights into which variables are important for a given problem (with the specific meaning of 'importance' depending on the measure used). A more traditional approach for this type of question would be to fit simple models to data, like linear regression, and to examine the coefficients of the fit. This approach is ubiquitous in science. Its basic drawback, however, is that one needs to assume a model, which may be inaccurate. For example, if the model is assumed to be y = mx+b, but the true relationship is y = cos x, then the value of m (the "interaction between x and y") will be 0 despite there being a strong relationship between x and y. The ML approach, on the other hand, seeks to maximize predictive accuracy and in doing so does not need to assume a simple functional form. This has the advantage that we can evaluate a variable's importance even when the relationship between inputs and outputs is unknown and possibly nonlinear. Plus, by bootstrapping, we can even find the confidence interval for their importance values. Machine learning combined with feature selection approaches can be universally applied to problems regardless of whether we know the underlying relationship. Determining the important features can also help us to construct simpler models. Rather than using many inputs for a model, we can only use the important features as inputs. For example, determining which morphological features of neurons are most predictive of cell type can lead us to build more accurate generative models of morphologies (that are based on the most predictive features). Accurately determining the importance of features within ML algorithms is thus also beneficial for creating simpler models. This approach is designed to examine the same variables that serve as raw inputs into the ML model. Often the "features" we seek are different than the raw inputs. In vision, for example, the raw input variables may be as simple as single pixels, and the approach of Role 2 would then simply isolate the most relevant pixels. However, as we review below, many problems in neuroscience follow the format where the input variables are the variables of interest. Neural Activity / Function Neuroscience has a long history of building encoding models, which aim to predict neural activity (e.g., spikes in an individual neuron, or BOLD signal in an fMRI voxel) based on variables in the outside world. This is a common approach to identifying the "role" of a brain area. The building of encoding models is a regression problem (from external variables to activity), and its purpose is more akin to feature importance than purely predictive power. This problem is an open invitation to use ML methods in combination with methods of determining feature importance. Machine learning would not be necessary for encoding models if simpler methods were as accurate at describing neural activity. However, this is usually not the case. For example, we recently showed that XGBoost and ensemble methods led to significant performance improvements on datasets from motor cortex, somatosensory cortex, and hippocampus (Benjamin et al. 2018)​. These improvements were relative to Generalized Linear Models, which are ubiquitous in computational neuroscience. Others have also shown predictive improvements in other areas and modalities using methods such as XGBoost ​(Viejo, Cortier, and Peyrache 2018) and deep learning ​(McIntosh et al. 2016; Klindt et al. 2017; Agrawal et al. 2014)​. These instances serve as warnings that although simple models may appear interpretable, they may be missing important aspects of how external variables relate to neural function. Having improved encoding performance can more generally allow researchers to understand which covariates are predictive of neural activity. This generalizes the already common approach of adding additional variables to simple models and observing the increase in performance (e.g., ​(Stringer et al. 2018)​). For example, research on building encoding models of ​(Viejo, Cortier, and Peyrache 2018) looked at the head-direction neurons using XGBoost relative contribution of the different covariates (such as the direction of the head) within the encoding model. This allowed determining how the covariates mattered, without assuming the form of the relationship. The reverse problem, "what information can be read-out from activity from this brain area" can also answer questions about information content and the role of specific brain areas or cell types. ​For example, researchers used decoding methods to compare how predictive neural populations in parietal and prefrontal cortices are about task-relevant variables, over the course of a decision-making task ​(Sarma et al. 2016)​. As another example, we have compared decoding results from motor cortex from different task conditions to determine how uncertainty in the brain relates to different behavioral uncertainty ​(Dekleva et al. 2016; Glaser et al. 2018)​. The choice of decoding method has a large impact on performance. We have recently done a thorough test of different ML methods on datasets from motor cortex, somatosensory cortex, and hippocampus, and have shown that modern ML methods, such as neural networks and ensemble methods lead to increased decoding accuracy ​(Glaser et al. 2017)​. More accurate decoding can increase our understanding of how much information is contained in a neural population about another variable, such as a decision, movement, or location. Neuroscience researchers often want to determine which variables matter for behavior, so that they can relate these variables to neural activity. We can apply ML to find what variables are predictive of behavior, without assuming the form of the relationship. For example, researchers have aimed to determine which visual features predict where we look next. This is a useful step in determining the neural basis of gaze control ​(Ramkumar et al. 2016)​. Traditionally, hand-designed visual features have been used to predict where we look next ​(Itti and Koch 2001)​, but recently researchers have more accurately predicted fixation locations using deep learning ​(Kümmerer, Theis, and Bethge 2014)​. As another example, researchers have studied how features in the environment predict the songs produced by male Drosophila during courtship ​(Coen et al. 2016)​. Using a generalized linear model with a sparsity prior, this study found that the distance between the male and female was the strongest predictor. This allowed the researchers to then investigate the neural pathway that was responsible for distance instance, modulating song amplitude. More accurate behavioral models can allow researchers to better investigate the relationship between neural activity and behavior. In medicine, it is important to understand the underlying factors that are predictive of disease. This has be done by finding the importance of neuroimaging features in traditional classification techniques (e.g., determining which functional connectivity measures are predictive of Alzheimer's disease in a logistic regression classifier ​(Challis et al. 2015)​). More recently, studies have used a variety of methods to look inside deep learning classifiers to determine the important features (e.g., determining fMRI connectivity relationships that are predictive of ADHD (Deshpande et al. 2015) and schizophrenia ​(Kim et al. 2016)​). This approach has also had success in animal models. For in a mouse model of depression, researchers determined which features (power and coherence) of neural activity in prefrontal cortex and limbic areas were predictive of pathological behavior ​(Hultman et al. 2016)​. They were then able to use this information to design a neural stimulation paradigm to restore normal behavior. By using machine learning in each of these applications, the researchers were able to test if their variables predicted disease without having to assume the form of that relationship. Neuroanatomy / Structure Just as with neural activity, machine learning can help researchers better understand how neuroanatomical features across the brain are predictive of disease. A general approach is to construct an ML classifier to determine whether a subject has the disease, and then look at the importance of the features (e.g., brain areas or connections) in that classifier. In one example, researchers trained an SVM classifier to predict depression based on graph-theory based features derived from diffusion-weighted imaging, and then looked at the importance of those features ​(Sacchet et al. 2015)​. In another example, researchers trained a random forest classifier to predict Alzheimer's disease from structural MRI and then determined which brain areas were the most predictive features in this classifier ​(Lebedev et al. 2014)​. Another general approach is to compare classification models that are constructed using different features. For example, the previously mentioned paper ​(Lebedev et al. 2014) also compared classifiers constructed with different feature sets, e.g., one using cortical thickness measures and one using volumetric measures. There are thus multiple ways in which ML can inform us about the predictive relationship between neuroanatomic features and neurological disease. Neurons have complicated shapes with varying biological structure that vary widely across brain regions and across species ​(Zeng and Sanes 2017)​. Many approaches have been proposed to classify neurons: electrophysiology ​(Teeter et al. 2018)​, morphology ​(Vasques et al. 2016)​, genetics or transcriptomics ​(Nelson, Sugino, and Hempel 2006)​, and synaptic connectivity (Jonas and Kording 2015)​. Machine learning can help with this endeavor ​(Armañanzas and Ascoli 2015; Vasques et al. 2016)​. The cell types can be labeled based on
1604.00091
1
1604
2016-04-01T01:02:30
Learning warps object representations in the ventral temporal cortex
[ "q-bio.NC" ]
The human ventral temporal cortex (VTC) plays a critical role in object recognition. Although it is well established that visual experience shapes VTC object representations, the impact of semantic and contextual learning is unclear. In this study, we tracked changes in representations of novel visual objects that emerged after learning meaningful information about each object. Over multiple training sessions, participants learned to associate semantic features (e.g. made of wood, floats) and spatial contextual associations (e.g. found in gardens) with novel objects. Functional magnetic resonance imaging was used to examine VTC activity for objects before and after learning. Multivariate pattern similarity analyses revealed that, after learning, VTC activity patterns carried information about the learned contextual associations of the objects, such that objects with contextual associations exhibited higher pattern similarity after learning. Further, these learning-induced increases in pattern information about contextual associations were correlated with reductions in pattern information about the objects visual features. In a second experiment, we validated that these contextual effects translated to real-life objects. Our findings demonstrate that visual object representations in VTC are shaped by the knowledge we have about objects, and show that object representations can flexibly adapt as a consequence of learning with the changes related to the specific kind of newly acquired information.
q-bio.NC
q-bio
Learning warps object representations in the ventral temporal cortex Alex Clarke1,2, Philip J. Pell1, Charan Ranganath2, Lorraine K. Tyler1 1Centre for Speech, Language and the Brain, Department of Psychology, University of Cambridge, UK. CB2 3EB. 2Center for Neuroscience, University of California Davis, Davis CA, 95618. Abstract The human ventral temporal cortex (VTC) plays a critical role in object recognition. Although it is well established that visual experience shapes VTC object representations, the impact of semantic and contextual learning is unclear. In this study, we tracked changes in representations of novel visual objects that emerged after learning meaningful information about each object. Over multiple training sessions, participants learned to associate semantic features (e.g. "made of wood", "floats") and spatial contextual associations (e.g. "found in gardens") with novel objects. Functional magnetic resonance imaging was used to examine VTC activity for objects before and after learning. Multivariate pattern similarity analyses revealed that, after learning, VTC activity patterns carried information about the learned contextual associations of the objects, such that objects with contextual associations exhibited higher pattern similarity after learning. Further, these learning-induced increases in pattern information about contextual associations were correlated with reductions in pattern information about the object's visual features. In a second experiment, we validated that these contextual effects translated to real-life objects. Our findings demonstrate that visual object representations in VTC are shaped by the knowledge we have about objects, and show that object representations can flexibly adapt as a consequence of learning with the changes related to the specific kind of newly acquired information. 1 Introduction The ventral temporal cortex (VTC) is crucial for object recognition (Clarke & Tyler, 2014; Grill-Spector et al., 1998; Kravitz, Saleem, Baker, Ungerleider, & Mishkin, 2013; Martin, 2007; Ungerleider & Mishkin, 1982) and many studies have demonstrated that visual experience can shape object representations in human VTC (see Kourtzi & Connor, 2011; Op de Beeck & Baker, 2010). In daily life, objects are not only processed according to their visual appearance, but also according to their meaning. Little is known, however, about how visual object representations change as they transition from being meaningless to meaningful. In order to understand how learning about meaning changes object representations, it is important to distinguish between different dimensions of meaning that could be learned. For instance, one can learn about intrinsic attributes (e.g. has ears, made of metal, floats) that determine the function or significance of an object, or about spatial contextual associations (e.g. found in the zoo) that enable objects to be situated in the world. Previous studies, in which participants learned semantic features for meaningless objects have shown learning-related increases in brain responses (Bellebaum et al., 2013; James & Gauthier, 2003, 2004; Moore, Cohen, & Ranganath, 2006; Skipper, Ross, & Olson, 2011; Vuilleumier, Henson, Driver, & Dolan, 2002; Weisberg, van Turennout, & Martin, 2007). These studies show that learning object meaning can change how objects are represented in VTC, but they do not address the central issue of how specific aspects of meaning drive changes in the neural representation of objects. The aim of the present study was to test if there is a relationship between the specific type of information people learned and how object representations changed. We used functional magnetic resonance imaging (fMRI) to examine how representations of pre-experimentally novel objects are modified as they become meaningful. We examined the effect of learning two aspects of meaning – an object's semantic category and its contextual association, both linked to regions of the VTC such as the posterior fusiform gyrus (Chao, Haxby, & Martin, 1999; Clarke & Tyler, 2014; Cox & Savoy, 2003; Huth, Nishimoto, Vu, & Gallant, 2012), parahipocampal cortex (PHC), and extending to the retrosplenial cortex (RSC; Aminoff, Gronau, & Bar, 2007; Bar & Aminoff, 2003; Bar, Aminoff, & Schacter, 2008; Stansbury, Naselaris, & Gallant, 2013). Further, we tracked how learning about different aspects of meaning impacts on visual shape-based representations. Participants learned a name and four semantic features for each of 12 novel objects, that created a semantic category structure (three different categories based on overlap of features), while objects either did or did not have a contextual feature. We used multivariate representational similarity analysis (RSA; Kriegeskorte, Mur, & Bandettini, 2008; Nili et al., 2014) to determine how learning these two aspects of meaning - semantic category and contextual associations- effects neural similarity spaces in VTC. We predict that if learning meaningful information drives changes in neural similarity spaces, then objects from the same semantic category, or those associated with a context, will show more similar activation patterns in VTC than objects not sharing that property. Within the VTC, we predict semantic category effects to be most prominent in the fusiform, while contextual association effects are predicted 2 to be more widespread including the RSC, PHC, lingual and fusiform. Finally, the visual shape similarity of the objects is predicted to relate to activation patterns in early visual cortex possibly extending into the VTC, and allows us to test for a relationship between visual form information and the newly learned semantic information. To assess the generalizability of our results, we ran a second experiment to examine how semantic category and contextual associations influence VTC activity patterns during processing of real-world objects. Based on previous studies we would expect category and contextual effects for real objects, and their inclusion here not only demonstrates the generalisability of any learning effects, but importantly allows for comparisons of the distribution of effects for real and novel objects which can help establish whether effects of our learning paradigm lead to the kinds of representational changes that mirror the long-term learned representations that are present for real objects. Methods Experiment 1: Novel objects Overview An overview of the experimental sessions can be seen in Table 1. The experiment consisted of two identical fMRI sessions (mean time between scans was 26 days; range 21-28 days) and four behavioural learning/testing sessions during which the semantic feature information was learned for 12 novel objects. During the fMRI sessions, participants performed a simple visual task that could be performed on the objects both before and after learning, and all behavioural sessions were conducted between the two fMRI sessions. The behavioural sessions were completed within one week, with behavioural session 3 occurring between 48 and 72 hours prior to the second fMRI session, and behavioural session 4 occurring just prior to the second scanning session. Participants Twelve healthy participants (6 males, 6 females) completed all MRI and behavioural sessions. All had normal or corrected to normal vision and were right handed. The average age was 20.7 years (range 19- 23 years). All participants provided informed consent, and the study was approved by the Cambridge Research Ethics Committee. One participant was excluded due to excessive head movement during both scanning sessions, and was excluded from all analyses (leaving eleven participants). Stimuli A total of 24 non-real, novel objects were used in the study - 12 were assigned to the learning condition and 12 to the exposure condition. The novel objects were "fribbles" (downloaded from http://wiki.cnbc.cmu.edu/Novel_Objects), and have a main body and four appendages. Eight fribbles were selected from three different visual "species", where a species is defined by a common main body. Within each species, two fribbles were selected from each of four "families", where members of a family share the main body and the appendages have a variable degree of overlap, and fribbles from the same 3 species but different families share a main body but have no appendages in common. This structure provides a variable degree of visual similarity across the novel objects based on the main body and appendages (visual features). Half of the fribbles were used for the learning condition and half for the exposure condition. In each condition, and for all species, the amount of visual feature overlap was matched to ensure that effects were due to the semantic, rather than the visual, properties of the objects. All fribbles were displayed as grey-scale images in the centre of a white background. Pretesting was used to ensure that the 24 fribbles did not have a strong resemblance to familiar object categories (animals, plant life, tools, vehicles, living, nonliving). Eight items from the exposure set were inverted based on pretests so they no longer showed a resemblance to familiar object categories. Figure 1. Objects in the learning condition, showing their names and semantic features. Each row shows a different semantic group, where each member shares 2 or 3 features. Each semantic group is composed of objects from two different visual groups. 4 The 12 fribbles from the learning condition were each assigned a name and four semantic features (e.g. made of metal; Figure 1) forming three semantic groups based on semantic feature overlap. Each semantic group contained four fribbles that were drawn from two different visual species. All items within a semantic group had two semantic features in common that were not present for items in other semantic groups (the shared features of that group i.e. Made of metal & Conducts electricity for the four fribbles on the top row of Figure 1). These shared features provide the basis for semantic similarity within each group and the basis for category organisation. Further, one semantic feature was shared between two items in each group (less shared features; e.g. Is durable is present in two fribbles on the top row of Figure 1) and all other semantic features were unique, with a total of 27 features. Twenty six of the semantic features were selected from the McRae et al., (2005) feature production norms, and one feature was chosen by the experimenters. All semantic features were non-visual (i.e., not describing visual object form) and synonyms were avoided to ensure features were easily discriminable. Features that signify animacy (e.g., eats) or specify the function of familiar items (e.g., used for holding liquids) were also avoided to increase the plausibility of the features for novel objects. The two shared semantic features were a material (e.g., made of wood) and a well known property of that material (e.g., burns). Two of the unique semantic features in each semantic group were contextual features (e.g. Found in...). The remaining semantic features were sensory (e.g., smells bad), hidden-visual (e.g., yellow on the inside) or general object properties (e.g. has layers, is lightweight). This selection of the fribbles and assignment of semantic features allows us to test for high-level visual, semantic feature and contextual similarities in the brain due to learning. Procedure Behavioural sessions Participants learned to associate semantic information for 12 novel objects over two learning/testing sessions on separate days (behavioural sessions 1 to 2) and two testing session (behavioural sessions 3 to 4). While learning was only given in the first two sessions, testing was conducted in all behavioural sessions to track learning rates over time. Learning sessions The 12 objects were learned in subsets of 4 items. Initially participants viewed information slides for each object in the subset. Slides contained an image of the object along with text containing its name and semantic features (Figure 2a). Participants were instructed they would be tested on their knowledge of the text and were asked to read each slide carefully. After viewing the information slides, participants answered a series of questions about the objects and their features (Figure 2b). Questions were presented in three phases; Phase 1 - associating objects with the shared features. Phase 2 - associating objects with less shared and unique features. Phase 3 - associating names with the objects/semantic features. Questions followed the general form of presenting a target and two response options. The nature of the targets and response options varied across trials and phases but could be an object image, name, single feature or pair of features. The response options remained on screen until a decision was made, and feedback was given on each trial to reinforce learning. Information slides for each object 5 were intermixed within the question trials. Each phase was repeated if accuracy was below 80%. In total there were 252 question trials in behavioural session 1 and 216 question trials in behavioural session 2, where only phases 2 and 3 were given. Figure 2. Semantic learning and testing. a) Example information slide presented to participants in order to learn semantic information. b) Example question trial from phase 2 of learning. c) Example question from the forced-choice recognition task in the testing phase Testing Two methods were used to measure learning performance: a forced-choice recognition task and a feature-recall task. Forced-choice recognition tests were administered after learning in behavioural 6 sessions 1 and 2, and again during behavioural session 3 (where no learning took place). On each trial, four semantic features belonging to the same object were presented with three options (Figure 2c), that were either object images or names. The incorrect options were either from the same or different semantic groups. There were a total of 48 trials in the testing sessions. Semantic features and response options remained on screen until a response was made, and no feedback was given. The lowest accuracy across all three testing sessions was 79% (Table 2), with all participants scoring higher than 85% in the final testing session and 5/11 participants scoring 100%. Feature-recall tests were administered at the start of behavioural session 2, 3 and 4. Participants were given two sheets of paper containing the images of the 12 learned, novel objects, above five empty lines labelled: "Name", "Feature 1", "Feature 2", "Feature 3", and "Feature 4". Participants were asked to fill in as much information as possible within a 10 minute time limit. The ordering and location of each object on the answer sheets was randomised across the three tests, and was not ordered by semantic or visual groupings. The lowest accuracy in the final testing session (behavioural session 4) was 67%, while 7/11 participants scored 100% (Table 2). Overall, the testing scores indicate that the vast majority of participants learned the associations between the semantic features and the object images to a high level of success (9/11 scored at least 98% in the forced-choice testing, and 8/11 scored at least 96% in the free-recall). Exposure condition An additional 12 novel objects were included to act as a control condition for any influence of visual familiarity and to test for overall effects of learning meaning. These objects were not associated with semantic features, but participants were exposed to them using a one-back visual matching task that was performed during behavioural sessions 2 and 3. On each trial an object appeared for 600 ms and was followed by a 1 s central fixation cross. Participants were instructed to press the right button if the object on screen did not match the previous item or the left button if it did match the previous item. A 12-trial practice of the one-back task was followed by two 96-trial blocks. Within each block there were 8 repetitions of each item and each item was twice a 'match' target; giving a total of 24 match-trials in each block. Participants were given feedback on their performance at the end of each block. fMRI sessions Procedure Both pre and post-learning fMRI sessions used the same task and procedure. Participants performed a visual anomaly detection task where they had to detect when one of the objects' visual features was 'bleached out'. The task ensured participants paid close attention to the images, could perform the task both prior to and post-learning, and could be equally performed with objects from the learning and exposure sets. Two modified versions of each of the 24 objects was created for the anomaly detection task by increasing the brightness and contrast of one of the appendage features by 60% using Adobe Photoshop CS2. A different feature was modified in each version. One version was used in the first scanning session and the other in the second scanning session. 7 Each trial consisted of a centrally presented black fixation cross on a white background for 300 ms, followed by a picture lasting 700ms, then a blank white background between 2 s and 7 s. The participants' task was to press one button if the object on screen was an unmodified image or another button if the object was a modified image (where a feature was bleached out; see Fig M4 centre panel below). There were 15 repetitions of each image; 12 unmodified versions and three modified versions. Objects were presented in 15 blocks each containing a single presentation of all 24 object with four or five modified images in each block. The order of items within a block was random, and different for every block. The position of the four/five modified items in each block was random with the constraint that no more than two could occur in succession. The presentation and timing of stimuli was controlled using Eprime version 1.1 (Psychology Software Tools, Pittsburgh, PA). Scanning All scanning took place at the MRC Cognition and Brain Sciences Unit, Cambridge, in a Siemens 3-T Tim Trio MRI scanner (Siemens Medical Solutions, Camberley, UK). Two functional scans were collected in each scanning session using gradient-echo echoplanar imaging (EPI) sequences collecting 32 slices in descending order of 3 mm thickness and between slice gap of 0.75 mm and a resolution of 3 x 3 mm. The field-of-view was 192 x 192 mm, matrix size 64 x 64 with a TE of 30 ms, TR of 2 s and a flip angle of 78⁰. Each functional scan lasted approximately 12 minutes with a short break half-way through. Prior to functional scanning, a high-resolution structural MRI image was collected using an MPRAGE sequence with 1 mm isotropic resolution. fMRI preprocessing Data from the two scanning sessions were preprocessed independently. Functional images were slice- time corrected, spatially realigned and smoothed using a 4 mm Gaussian kernel in SPM8 (Wellcome Institute of Cognitive Neurology, London, UK). These unnormalised images were analysed for each participant with the general linear model to creating a single beta image for each object based on the 12 repetitions of the unmodified images. In addition to the 24 novel object predictors in the GLM, predictors were included to capture effects associated with the modified objects, slow trends using 11 regressors for the first block of functional scans and 12 for the second block based on the basis functions of a discrete cosine transform (minimum frequency = 1/128 Hz), six head motion regressors for each session along with their first derivatives, and a global mean predictor for each scanning session. We also included a separate regressor for each fast-motion event, where a fast motion event was defined as motion greater than 0.7 mm/TR and detected using ArtRepair software (Mazaika, Hoeft, Glover, & Reiss, 2009). The resulting 24 beta images for the unmodified objects were converted to t-images prior to the RSA analyses. RSA analysis We first used a searchlight mapping approach (Kriegeskorte, Goebel, & Bandettini, 2006) to test for either consistent effects over sessions, or learning-induced changes to how the novel objects were represented in the pre- and post-learning fMRI sessions. Searchlight analyses were followed by analyses of anatomically defined ROIs that have been previously linked to semantic category and contextual effects, and was performed as activation patterns may encode information at the spatial scale of 8 anatomical regions which searchlight analyses may be less sensitive to due to their smaller spatial neighbourhood (Etzel, Zacks, & Braver, 2013). Candidate model RDMs We tested a visual model RDM and three other models capturing different semantic distinctions (Figure 3). The visual features model captures the amount of visual-feature overlap between pairs of objects, where a feature can be the main body or an appendage. This is a high-level model of visual similarity that captures visual shape information, rather than simply low-level visual information. Objects from the same fribble species will always share at least one feature (the main body), while objects from different fribble species always share no features. The meaning model tests for overall effects of learning meaning in contrast to the objects in the exposure condition. Here, objects that were learned about are predicted to cluster together to a greater degree that objects from the exposure condition, which have no associated meaning. The semantic category model captures the three semantic categories the objects belonged to, where members of the same semantic category share either 2 or 3 features, and no features were shared across semantic categories. The contextual model tests where activation patterns for the learned objects form two clusters according to whether they were associated with a specific context or have no contextual association (the correlation between the semantic category and contextual association RDMs was r = 0.19). Figure 3. Candidate model RDMs tested using representational similarity analysis. Searchlight Each candidate model RDM was tested against the observed activation patterns using a searchlight similarity analysis implemented in the RSA toolbox (Nili et al., 2014) and custom matlab functions, 9 before we tested for significant conjunctions or changes over sessions. At each voxel, object activation values from grey matter voxels within a spherical searchlight (radius 9 mm) were extracted to calculate distances between all items (using 1 - Pearson correlation) creating an object dissimilarity matrix based on that searchlight. This fMRI RDM was correlated with each candidate model RDM (using Spearman's rank correlation) and the resulting similarity values were Fisher transformed and mapped back to the voxel at the centre of the searchlight. Similarity maps for each model RDM and each participant were normalised to the MNI template space and spatially smoothed using a 6 mm FWHM Gaussian kernel. The similarity maps for each participant were then entered into a group-level random effects (RFX) analyses using SPM8. We used a paired samples t test design testing for first, common RSA effects across both sessions using a conjunction analysis (the conjunction null hypothesis) for positive effects, and second, increased RSA effects in the post-learning session compared to the pre-learning session. Given our a priori expectation that effects will be in the ventral processing stream, these analyses were constrained to a ventral stream mask (including the RSC) produced by combining bilateral regions from the Harvard-Oxford brain atlas (occipital pole, intracalcarine cortex, cuneal cortex, lateral occipital inferior, fusiform occipital, fusiform temporooccipital, lingual gyrus, parahippocampal posterior) and the RSC (BA29 & 30). Results are reported using a voxelwise threshold of p < 0.005 and a cluster extent of p < 0.05 (FWE) correcting for multiple comparisons. An anatomical ROI analysis was performed to test for effects of the candidate model RDMs in targeted VTC regions that have been previously implicated in semantic category and contextual processing. This was done to test for distributed regional effects that may only be present in activation patterns at the spatial scale of regions, rather than the smaller searchlights. This would be the case if information was distributed across voxels at a spatial scale larger than the searchlights. A number of ROIs were specified to cover the VTC, specifically the fusiform (fusiform temporooccipital), lingual (lingual gyrus) and parahippocampal (parahippocampal posterior). ROIs were defined from the Harvard-Oxford brain atlas (Desikan et al., 2006), and the retrosplenial cortex was also included and defined as BA29 and BA 30. For each ROI, object activation values were extracted for all voxels in the ROI to calculate distances between the 24 items (using 1 - Pearson correlation). The fMRI ROI RDMs were correlated with the candidate model RDM (using Spearman's rank correlation) and the resulting similarity values were Fisher transformed. RFX analyses were performed using one-tailed one-sampled t tests against zero when testing for significant positive effects of each model RDM, and paired-samples t-tests when testing for significant differences between fMRI sessions or model RDMs. ROI analysis Experiment 2: Real objects Participants Sixteen participants took part in the study (6 male, 10 female). All were right-handed and were aged between 19 and 29 (mean 23 years). All participants had normal or corrected to normal vision, and gave informed consent. The study was approved by the Cambridge Research Ethics Committee. 10 Stimuli A total of 145 real objects were used, where 131 of these were from one of six object categories (34 animals, 15 fruit, 21 vegetables, 27 tools, 18 vehicles, 16 musical instruments) and 14 additional objects that did not adhere to a clear category (and were not included in our analyses). Isolated coloured objects were shown in the centre of a white background, and normalised to a maximum visual angle of 7.5°. All objects were chosen to depict concepts from an anglicised version of the McRae production norms (McRae et al., 2005; Taylor, Devereux, Acres, Randall, & Tyler, 2012) from which semantic feature information could be obtained to construct the meaningful model RDMs. Procedure Participants performed an overt basic-level naming task. Each trial consisted of a fixation cross lasting 500 ms, before an object for 500 ms followed by a blank screen lasting between 3 and 11 seconds. All objects were repeated 6 times across 6 different blocks. The object presentation order for each block was randomised for each participant, although a constant category order was maintained ensuring an even distribution of object category across the block. The presentation and timing of stimuli were controlled with Eprime version 1 (Psychology Software Tools, Pittsburgh, PA, USA), and naming accuracy was recorded by the experimenter during acquisition. fMRI acquisition Participants were scanned at the MRC Cognition and Brain Sciences Unit, Cambridge, in a Siemens 3-T Tim Trio MRI scanner (Siemens Medical Solutions, Camberley, UK). There were 3 functional scanning sessions using gradient-echo echoplanar imaging (EPI) sequences collecting 32 slices in descending order of 3 mm thickness and between slice gap of 0.75 mm and a resolution of 3 x 3 mm. The field-of-view was 192 x 192 mm, matrix size 64 x 64 with a TR of 2 seconds, TE of 30 ms and a flip angle of 78⁰. Each functional session lasted approximately 9-10 minutes, containing two object blocks. Prior to functional scanning, a high-resolution structural MRI image was collected using an MPRAGE sequence with 1 mm isotropic resolution. fMRI preprocessing Functional images were slice-time corrected, spatially realigned and smoothed using a 4 mm Gaussian kernel in SPM8 (Wellcome Institute of Cognitive Neurology, London, UK). These unnormalised images were analysed for each participant with the general linear model to create a single beta image for each object based on the 6 repetitions. In addition to the 145 object predictors, predictors were included to capture slow trends using 18 regressors for each session based on the basis functions of a discrete cosine transform (minimum frequency = 1/128 Hz), six head motion regressors for each session and a global mean predictor for each scanning session. The resulting 145 beta images were converted to t- images prior to the RSA analyses. Only objects named correctly on all six repetitions (86%, SE = 1.53%) were included in further analyses. RSA analysis An RSA analysis was performed for the real objects using both searchlight mapping and an anatomical ROI analysis. The procedures were identical to those used in experiment 1, with the exception that here 11 the searchlight RFX analysis was conducted using a one-sampled t-test testing for positive effects of the models with no a priori voxel restrictions. We tested for RSA effects of a contextual association RDM and a semantic category RDM (note that the high-level visual features and meaning models were not tested, as they can not be defined for our real objects in the same manner as they are defined for the novel objects). The contextual association model RDM was defined for real objects in the same manner as we defined it for novel objects. Using the McRae et al. (2005) production norm data to extract semantic features for our real objects, we can determine if an object has contextual associations or not. Features that indicated an object has strong contextual associations took the form; associated with X, found in/near/on X, lives in X, used at/on X, where X signified a specific location (e.g. deserts, school, Spain). Based on this criterion, 44 out of 131 objects were determined to have spatial contextual associations. The contextual association model RDM for real objects was constructed in the same way as for novel objects, where objects with contextual associations will form one cluster, and objects without contextual associations are form a second cluster. The semantic category RDM captures the category structure of the 131 objects across 6 superordinate categories (animals, fruit, vegetables, tools, vehicles, musical instruments) and indicates there is more within-category similarity than between-category similarity. The correlation between the semantic category and contextual association RDMs was r = 0.09. Results Experiment 1: Novel objects The goal of this study was to determine whether learning-induced changes in object representations are governed by the semantic category and contextual association information people learn, and whether this impacts on visual form representations. The similarity relations between the objects, as embodied in the model RDMs, capture the visual and meaningful (learned) similarity spaces that can be tested against brain activation patterns to track changes in the information represented due to learning. Here, we used RSA to uncover statistical correspondences between the visual and meaningful (i.e. learned) similarity structures (Figure 3) and activation pattern similarities across the different objects. We first explored whether effects of our candidate model RDMs showed significant learning-induced changes, or consistent effects across sessions, in their relationship to activation patterns. To do this we directly compared the RSA searchlight maps from the pre- and post-learning scanning sessions, where the pre- learning session acts as a baseline for how the brain responds to meaningless objects. We then present an anatomical ROI analysis to look for more spatially distributed effects that extend beyond the size of our searchlights. Using searchlight RSA, the visual features model was found to show consistent significant effects across both sessions in the occipital lobe (peak MNI coordinate: 21, -88, -1; Figure 4a), showing that searchlights within the visual cortex responded to the same images in a similar manner both before and after learning. Multidimensional scaling (MDS) further illustrated that the regional activation patterns 12 clustered by visual form similarity, with objects from the same visual groups (shown by different colours) falling closer together in this 2-dimensional space than objects from different visual groups. Testing for increased RSA effects after learning compared to before, we found significant effects relating to the contextual association model. Representational changes were seen in the right VTC with two foci – one on the fusiform gyrus (change in Spearman's rho = 0.08; peak MNI coordinate: 39, -34, -24) and one spanning the collateral sulcus (change in Spearman's rho = 0.04; peak MNI coordinate: 24, -52, -16; Figure 4b). MDS plots further showed how activation patterns in both these areas dissociate along the contextual association dimension of the stimuli. No effects were seen for the other meaning related models. These results show that activation patterns in the VTC have been warped through learning, whereby the same stimuli are represented in a different manner after learning in comparison to before, while object activation patterns in the early visual regions remained more stable in that they were significantly related to the visual feature information in both sessions, that were not significantly different to one another (Note that equivalent results are found by inspecting pre- and post- learning searchlight maps separately). Figure 4. RSA searchlight effects for novel objects. a) Searchlights showing common significant effects across both sessions for the visual features model. MDS plot derived from object pattern similarities averaged over sessions with each visual group shown in a different colour. b) Searchlights showing significant increases for the contextual associations model in the post-learning session compared to the pre-learning session. MDS plots derived from object pattern similarities in the post-learning session for the two foci, where objects with contextual associations shown in blue and without contextual associations in red (exposure objects not shown for clarity). Both images are voxelwise p < 0.005, cluster p < 0.05. 13 While our searchlight analysis tested for effects in local activation patterns (i.e. searchlights), representational information could be present at the larger scale of anatomical regions which will not always be captured by smaller searchlights. This would be the case if information was distributed across voxels at a spatial scale larger than the searchlights. Therefore we also performed an analysis testing the candidate model RDMs within anatomically defined regions across the posterior ventral temporal lobe, including the RSC, that have been implicated in semantic category and contextual processing - specifically bilateral posterior fusiform, lingual, PHC and RSC (Figure 5). Figure 5. RSA ROI analyses showing pre- and post-learning effects for the 4 model RDMs in the 4 anatomical ROIs. Pre-learning effects shown in light grey, post-learning effects in dark grey. Asterisks show significant effects of the models (**p<0.01,*p<0.05) and significant changes over sessions shown by a horizontal bar. PHC = parahippocampal cortex and RSC = retrosplenial cortex. Error bars show standard errors. Confirming the results from our searchlight analysis, we observed significant learning-induced increases in contextual association information in the fusiform in the post-learning session compared to the pre- learning session (t(10) = 3.37, p = 0.004). Further, the fusiform region showed significant effects of the contextual association model during the post-learning session (t(10) = 5.10, p = 0.0002), which were absent in the pre-learning session (t(10) = 0.6). We also observed marginally significant learning-induced increases for the contextual association model in the RSC (t(10) = 1.67, p = 0.064), which was significant for the post-learning session (t(10) = 1.96, p = 0.039) but not in the pre-learning session (t(10) = 0.05). Finally, the lingual region showed significant effects for the contextual association model in the post- learning session (t(10) = 1.92, p = 0.042), and not in the pre-learning session (t(10) = 0.4), although no significant change was seen between the two sessions (t(10) = 1.1). Turning to the visual features model, a significant effect was seen in the fusiform for the pre-learning activation patterns (t(10) = 1.92, p = 0.042), that was reduced in the post-learning session and no longer significant (t(10) = 0.6), although cross session comparisons revealed no significant differences between 14 the two sessions (t(10) = 0.85). The visual feature model also showed effects in the lingual region after learning (t(10) = 3.02, p = 0.006), which showed marginal effects in the pre-learning session (t(10) = 1.62, p = 0.068) that was not significantly different across the two scanning sessions (t(10) = 0.2) suggesting the lingual region reflected relatively stable object representations, similar to our searchlight results. Finally, we note that the other two meaning-related models – the meaning and semantic category RDMs, both showed significant effects in the post-learning session only (meaning: t(10) = 2.41, p = 0.018, semantic category: t(10) = 2.33, p = 0.021). These effects were absent in both the cross-session searchlight comparisons and in individual searchlight effects for the post-learning session (not shown) suggesting that information relating to general meaningfulness, and to the learned semantic categories the objects belong to, may be coded in more distributed patterns than the searchlight mapping was sensitive to. The results from the anatomical ROI analyses confirm what was observed in the our searchlight analyses - learning about novel objects induced representational changes in VTC, most prominently in the fusiform gyrus, where patterns reflected the learned contextual associations of objects. Within the anatomical ROIs we also saw suggestions of a more general meaning-related effect conferred through the meaning and semantic category models. Moreover, after learning, visual similarity effects in the fusiform are reduced and learning-induced contextual similarities emerge. Testing the relationship between representational changes in the fusiform over the sessions showed a significantly greater change in RSA effects for the contextual model compared to both the visual feature model (t(10) = 2.2, p = 0.05) and the semantic category model (t(10) = 2.2, p = 0.05). Moreover, the reduction in the visual feature model effect was significantly correlated with the increased effect for the contextual association model (R-sq = 0.49, r = -0.7, p = 0.0085; Figure 6) indicating a representational shift in the kind of information that was represented in the fusiform as a consequence of learning. Figure 6. Relationship of representational changes between pre- and post-learning sessions for visual feature and contextual association model RDMs. Experiment 2: Real objects In Experiment 1 we found that learning about contextual associations for pre-experimentally novel objects changed voxel pattern similarity information in VTC. However, in real life, contextual associations are learned by encountering objects in particular contexts, whereas for our novel objects associations were learned through reading text. Accordingly, in Experiment 2, we use RSA to test for 15 effects of a contextual association model RDM for real objects in a previously published dataset (Clarke & Tyler, 2014). This not only allows us test the generalisability of our learning effects, but allows us to compare the distribution of effects for real and novel objects. Further we test for effects of a semantic category model, also known to show effects in the VTC, as a means of comparison, using both searchlight mapping and the anatomical ROIs within the VTC. Searchlight RSA showed effects for the contextual association model primarily in bilateral posterior VTC including the fusiform, lingual, and parahippocampal cortices, and also including the calcarine and inferior occipital gyrus (peak MNI coordinate: -9, -91, 10; Figure 7a). The results echo those previously reported for the contextual associations of real world objects (e.g. Stansbury et al., 2013), validating the manner in which we define our contextual model, and partly overlapping with our effects for novel objects in the collateral sulcus. The semantic category effects were widespread throughout VTC (peak MNI coordinate: -27, -26, 59), similar to previously reported effects (e.g. Connolly et al., 2012; Huth et al., 2012) although we would also highlight that there are well known confounds between an object's semantic category and its visual properties, a factor that we controlled in our novel objects experiment. Figure 7. RSA effects of the contextual association and semantic category models using real objects. a) Searchlight results, shown thresholded at voxelwise p < 0.005, cluster p < 0.05. b) RSA effects for the anatomical ROIs. Dark grey bars show effects for the contextual association model and light grey for the semantic category model. Asterisks show significant effects of the models (**p<0.01,*p<0.05), with significant changes over sessions shown by a horizontal bar. PHC = parahippocampal cortex, RSC = retrosplenial cortex, VTC = ventral temporal cortex. Error bars show standard errors. Within the anatomical ROIs, we found significant effects of the contextual association model in the fusiform (t(15) = 4.42, p = 0.0002), lingual (t(15) = 3.34, p = 0.0022) and PHC (t(15) = 2.13, p = 0.025) confirming our searchlight results are also present at this larger spatial scale (Figure 7b). The semantic category model showed significant effects in the fusiform (t(15) = 8.61, p < 0.0001), lingual (t(15) = 4.13, p = 0.0004), PHC (t(15) = 2.44, p = 0.014), and the RSC (t(15) = 2.34, p = 0.017). Further, the ROI similarity matrix was more similar to the semantic category model than the contextual association model in the fusiform ROI (t(15) = 5.10, p = 0.0001), showing the inverse relationship to that found for novel objects. Overall, both searchlight and ROI analyses showed activation patterns to real objects show a significant 16 relationship to the contextual associations model with the strongest effects in the fusiform and lingual gyri, plus widespread semantic category effects across VTC that also peaked in the fusiform gyrus. Discussion The goal of the present study was to determine how learning different kinds of meaningful information affects the neural representation of objects in VTC. In Experiment 1, we examined the effect of learning two aspects of meaning: the semantic features of an object, and whether each object had a specific contextual association (i.e., that an object is found in a particular setting). We found that learning about the presence of contextual associations induced systematic changes in the similarity structure of corresponding object representations in VTC. Learning led to the development of two clusters in the similarity space, one consisting of objects with a strong contextual association (i.e. a 'Found in...' feature) and another consisting of objects that did not have a strong contextual association. In Experiment 2, we demonstrated a distinction between images of real-world objects that had prominent contextual associations and objects that did not elicit strong contextual associations. Collectively, the results suggest that learned contextual associations exert powerful influences on the neural mechanisms of object processing. We elaborate on this and related issues below. Contextual association effects in VTC Searchlight-based voxel pattern similarity analyses in Experiment 1 revealed that learning contextual associations affected object similarity spaces in VTC (Figure 4). Within VTC, there was a lateral peak in the fusiform gyrus and a more medial peak spanning the collateral sulcus where the similarity responses reflected the presence of learned contextual associations. The peak of the medial effect was posterior to the parahippocampal gyrus, peaking closely to reported scene selective responses (Aguirre, Zarahn, & D'Esposito, 1998). Anatomical ROI analyses converged with these results, showing contextual association effects for both novel and real objects in the fusiform, and changes for novel objects were also seen in RSC. Together, the findings demonstrate that learning about contextual associations causes significant representational changes in the VTC. Large areas of the VTC have been shown to contain information about natural scene categories and their contextual associations (Stansbury et al., 2013; Walther, Caddigan, Fei-Fei, & Beck, 2009), in addition to meaningful object information more generally (Huth et al., 2012; Mahon, Anzellotti, Schwarzbach, Zampini, & Caramazza, 2009; Tyler et al., 2013; Vuilleumier et al., 2002). The emergence of learned information suggests that representations in the VTC are malleable through short term learning, with this area strongly implemented in previous learning studies (Gauthier, Tarr, Aanderson, Skudlarski, & Gore, 1999; James & Gauthier, 2004; Kourtzi, Betts, Sarkheil, & Welchman, 2005; Moore et al., 2006; Op de Beeck, Baker, DiCarlo, & Kanwisher, 2006; Sigala & Logothetis, 2002; van der Linden, Murre, & van Turennout, 2008). Our results add to this assertion by showing that learning-induced changes are specific to the kind of meaning that was learned. 17 Whereas previous fMRI research claims the RSC processes contextual information, the PHC is also strongly implicated (Aminoff et al., 2007; Bar & Aminoff, 2003; Bar et al., 2008; Ranganath & Ritchey, 2012). In experiment 1, verbal learning about the context in which an object could be found had strong effects on pattern similarity relationships in RSC, but interestingly, no comparable effects were detected in PHC. In contrast experiment 2, using real-life objects, did show contextual association effects in the PHC. This differential pattern in the PHC could be attributed to how contextual information was learned in the cases of novel and real-life objects. In experiment 1, these contextual associations were learned verbally, but never visually experienced. This contrasts with our understanding of real-world objects, for which we predominantly acquire strong contextual associations through repeated visual exposure in specific contexts. These may suggest that visual experience is key for contextual association effects in the PHC, while contextual effects in the RSC were seen for novel objects that suggests a more abstracted role of linking objects to contexts that is not as tied to visual contexts as the PHC (Bar, 2004). Learning-induced informational warping in the fusiform Results from Experiment 1 demonstrated that learning meaningful information about novel objects was associated with changes in visual-form similarity relationships in the fusiform. Our anatomical ROI analysis showed that the change in the contextual association model effect over sessions was significantly greater than the change in the visual-features model effect, and critically, the reduction in the visual feature effect with learning was strongly correlated with the emergence of the contextual association effect. This result could reflect a dimensional modulation of the representational space (Folstein, Palmeri, & Gauthier, 2014) where there is a representational shift away from the pre-learning visually-based object information, to information about object meaning. This result must reflect a change in how these objects are represented in VTC, rather than a top-down effect, as the same visual task was used before and after learning that did not require specific access to object meaning. Our results further highlight how the VTC can flexibly code different kinds of object properties gained through experience (such as form and meaning) and suggests that meaning plays a critical role in shaping object representations. Beyond contextual association effects of meaning Beyond learned effects of contextual associations in the fusiform, we also found evidence for more general semantic effects after learning at the spatial scale of anatomical regions. Activation patterns in the fusiform ROI correlated with the meaning and semantic category RDMs after learning which may indicate more subtle changes along semantic dimensions, where members of the same semantic category show more similar activation patterns after learning. A number of explanations could underlie the more modest semantic category effects compared to the contextual effects for novel objects. One account would be that participants in Experiment 1 learned to associate objects with labels that did not correspond to pre-existing object categories (such as animals or tools). Categories defined by their shared features (e.g. one category would be composed of metal, electrically conducting things), as in Experiment 1, have little ecological relevance, and participants could not readily leverage pre-existing category representations to facilitate learning (Op de Beeck & Baker, 2010). 18 This may suggest that learning about object contexts and forming superordinate categories takes place over different timescales. The significantly stronger effects for the contextual model over the semantic category model in Experiment 1 in the fusiform could be a consequence of enhanced initial learning for the contextual associations, which would be beneficial as context provides additional semantic constraints. Ad hoc testing supports the notion that contextual information aids initial learning, as our participants showed higher behavioural feature-recall accuracy for objects with a contextual ('Found in...') feature following the first learning session (session 2: contextual mean feature-recall 66%, non- contextual mean feature-recall 54%, t(10) = 2.27, p = 0.046) with this advantage disappearing with further training sessions (session 3: contextual mean feature-recall 91%, non-contextual mean feature- recall 85%, t(10) = 2.12, p = 0.06; session 4: contextual mean feature-recall 94%, non-contextual mean feature-recall 92%, t(10) = 1.16, p = 0.27). The ability to form generalisations across items that share common features could require more extensive exposure to exemplars, especially when objects are novel and cannot readily fit into existing categories. Consistent with the possibility of different timescales of context and category learning, the contextual association effect was significantly greater than the semantic category effect in the fusiform for novel objects, whereas the inverse was found for real objects. One caveat is that, in the case of real objects, semantic category and visual properties are highly confounded, whereas this was not the case for novel objects in Experiment 1. Further research is therefore needed to clarify how contextual and category information interact in the formation of stable meaningful representations in the brain. In conclusion, the present results demonstrate that visual object representations change as a consequence of learning meaningful information for previously meaningless objects. Learning-induced changes in neural pattern similarity relationships tracked the specific kinds of information that participants learned about objects, showing how representations in the VTC can flexibly adapt to represent newly acquired meaningful information. Such changes were specific to the informational structure that was learned, further emphasising that object representations are shaped by the nature of the information we learn. Acknowledgements This project has received funding to LKT from the European Research Council (ERC) under the European Union's Horizon 2020 research and innovation programme (grant agreement No 669820), from the European Research Council under the European Community's Seventh Framework Programme (FP7/2007-2013)/ ERC Grant agreement n° 249640, and a Guggenheim Fellowship to CR. References Aguirre, G. K., Zarahn, E., & D'Esposito, M. (1998). An area within human ventral cortex sensitive to "building" stimuli: evidence and implications. Neuron, 21(2), 373-383. 19 Aminoff, E., Gronau, N., & Bar, M. (2007). The parahippocampal cortex mediates spatial and nonspatial associations. Cerebral Cortex, 17(7), 1493-1503. doi: 10.1093/cercor/bhl078 Bar, M. (2004). Visual objects in context. Nat Rev Neurosci, 5(8), 617-629. doi: 10.1038/nrn1476 Bar, M., & Aminoff, E. (2003). Cortical analysis of visual context. Neuron, 38(2), 347-358. Bar, M., Aminoff, E., & Schacter, D. L. (2008). Scenes unseen: the parahippocampal cortex intrinsically subserves contextual associations, not scenes or places per se. Journal of Neuroscience, 28(34), 8539-8544. doi: 10.1523/JNEUROSCI.0987-08.2008 Bellebaum, C., Tettamanti, M., Marchetta, E., Della Rosa, P., Rizzo, G., Daum, I., & Cappa, S. F. (2013). Neural representations of unfamiliar objects are modulated by sensorimotor experience. Cortex, 49(4), 1110-1125. doi: 10.1016/j.cortex.2012.03.023 Chao, L. L., Haxby, J. V., & Martin, A. (1999). Attribute-based neural substrates in temporal cortex for perceiving and knowing about objects. Nature Neuroscience, 2(10), 913-919. Clarke, A., & Tyler, L. K. (2014). Object-specific semantic coding in human perirhinal cortex. Journal of Neuroscience, 34(14), 4766-4775. Connolly, A. C., Guntupalli, J. S., Gors, J., Hanke, M., Halchenko, Y., Wu, Y., . . . Haxby, J. V. (2012). The representation of biological classes in the brain. The Journal of Neuroscience, 32(8), 2608-2618. Cox, D. D., & Savoy, R. L. (2003). Functional magnetic resonance imaging (fMRI) "brain reading": detecting and classifying distributed patterns of fMRI activity in human visual cortex. Neuroimage, 19(2 Pt 1), 261-270. doi: S1053811903000491 [pii] Desikan, R. S., Segonne, F., Fischl, B., Quinn, B. T., Dickerson, B. C., Blacker, D., . . . Killiany, R. J. (2006). An automated labeling system for subdividing the human cerebral cortex on MRI scans into gyral based DOI 10.1016/j.neuroimage.2006.01.021 Neuroimage, 968-980. interest. regions of 31(3), doi: Etzel, J. A., Zacks, J. M., & Braver, T. S. (2013). Searchlight analysis: promise, pitfalls, and potential. Neuroimage, 78, 261-269. doi: 10.1016/j.neuroimage.2013.03.041 Folstein, J. R., Palmeri, T. J., & Gauthier, I. (2014). Perceptual advantage for category-relevant perceptual Front Psychol, 5, 1394. doi: and motion. case of shape dimensions: 10.3389/fpsyg.2014.01394 the Gauthier, I., Tarr, M. J., Aanderson, A., Skudlarski, P., & Gore, J. C. (1999). Activation of the middle fusiform 'face area' increases with expertise in recognizing novel objects. Nature Neuroscience, 2, 568-573. Grill-Spector, K., Kushnir, T., Hendler, T., Edelman, S., Itzchak, Y., & Malach, R. (1998). A sequence of object processing stages revealed by fMRI in the human occipital lobe. Human Brain Mapping, 6, 316-328. Huth, A. G., Nishimoto, S., Vu, A. T., & Gallant, J. L. (2012). A continuous semantic space describes the representation of thousands of object and action categories across the human brain. Neuron, 76(6), 1210-1224. doi: 10.1016/j.neuron.2012.10.014 James, T. W., & Gauthier, I. (2003). Auditory and action semantic features activate sensory-specific perceptual brain regions. Current Biology, 13(20), 1792-1796. James, T. W., & Gauthier, I. (2004). Brain areas engaged during visual judgments by involuntary access to novel semantic information. Vision Research, 44(5), 429-439. 20 Kourtzi, Z., Betts, L. R., Sarkheil, P., & Welchman, A. E. (2005). Distributed neural plasticity for shape learning in the human visual cortex. PLoS Biol, 3(7), e204. doi: 10.1371/journal.pbio.0030204 Kourtzi, Z., & Connor, C. E. (2011). Neural representations for object perception: structure, category, and adaptive coding. Annual Review of Neuroscience, 34, 45-67. doi: 10.1146/annurev-neuro- 060909-153218 Kravitz, D. J., Saleem, K. S., Baker, C. I., Ungerleider, L. G., & Mishkin, M. (2013). The ventral visual pathway: an expanded neural framework for the processing of object quality. Trends Cogn Sci, 17(1), 26-49. doi: 10.1016/j.tics.2012.10.011 Kriegeskorte, N., Goebel, R., & Bandettini, P. (2006). Information-based functional brain mapping. Proceedings of the National Academy of Sciences of the United States of America, 103(10), 3863- 3868. Kriegeskorte, N., Mur, M., & Bandettini, P. (2008). Representational similarity analysis - connecting the branches of systems neuroscience. Front Syst Neurosci, 2, 4. doi: 10.3389/neuro.06.004.2008 Mahon, B. Z., Anzellotti, S., Schwarzbach, J., Zampini, M., & Caramazza, A. (2009). Category-specific organization in the human brain does not require visual experience. Neuron, 63(3), 397-405. Martin, A. (2007). The representation of object concepts in the brain. Annual Review of Psychology, 58, 25-45. doi: 10.1146/annurev.psych.57.102904.190143 Mazaika, P., Hoeft, F., Glover, G. H., & Reiss, A. L. (2009). Methods and Software for fMRI Analysis for Clinical Subjects. Paper presented at the Human Brain Mapping. McRae, K., Cree, G. S., Seidenberg, M. S., & McNorgan, C. (2005). Semantic feature production norms for a large set of living and nonliving things. Behavior Research Methods, 37, 547-559. Moore, C. D., Cohen, M. X., & Ranganath, C. (2006). Neural mechanisms of expert skills in visual working memory. Journal of Neuroscience, 26(43), 11187-11196. doi: 10.1523/JNEUROSCI.1873-06.2006 Nili, H., Wingfield, C., Walther, A., Su, L., Marslen-Wilson, W., & Kriegeskorte, N. (2014). A toolbox for doi: Comput Biol, e1003553. analysis. 10(4), PLoS representational similarity 10.1371/journal.pcbi.1003553 Op de Beeck, H. P., & Baker, C. I. (2010). The neural basis of visual object learning. Trends Cogn Sci, 14(1), 22-30. doi: 10.1016/j.tics.2009.11.002 Op de Beeck, H. P., Baker, C. I., DiCarlo, J. J., & Kanwisher, N. G. (2006). Discrimination training alters object representations in human extrastriate cortex. Journal of Neuroscience, 26(50), 13025- 13036. doi: 10.1523/JNEUROSCI.2481-06.2006 Ranganath, C., & Ritchey, M. (2012). Two cortical systems for memory-guided behaviour. Nat Rev Neurosci, 13(10), 713-726. doi: 10.1038/nrn3338 Sigala, N., & Logothetis, N. K. (2002). Visual categorization shapes feature selectivity in the primate temporal cortex. Nature, 415, 318-320. Skipper, L. M., Ross, L. A., & Olson, I. R. (2011). Sensory and semantic category subdivisions within the doi: Neuropsychologia, 3419-3429. 49(12), anterior 10.1016/j.neuropsychologia.2011.07.033 temporal lobes. 21 Stansbury, D. E., Naselaris, T., & Gallant, J. L. (2013). Natural scene statistics account for the representation of scene categories in human visual cortex. Neuron, 79(5), 1025-1034. doi: 10.1016/j.neuron.2013.06.034 Taylor, K. I., Devereux, B. J., Acres, K., Randall, B., & Tyler, L. K. (2012). Contrasting effects of feature- based statistics on the categorisation and identification of visual objects. . Cognition, 122(3), 363-374. Tyler, L. K., Chiu, S., Zhuang, J., Randall, B., Devereux, B. J., Wright, P., . . . Taylor, K. I. (2013). Objects and categories: Feature statistics and object processing in the ventral stream. Journal of Cognitive Neuroscience, 25(10), 1723-1735. Ungerleider, L., & Mishkin, M. (1982). Two cortical visual systems. In D. J. Ingle, M. A. Goodale, & R. J. W. Mansfiled (Eds.), Analysis of visual behavior (pp. 549-586). Cambridge, MA: MIT Press. van der Linden, M., Murre, J. M., & van Turennout, M. (2008). Birds of a feather flock together: experience-driven formation of visual object categories in human ventral temporal cortex. PLoS ONE, 3(12), e3995. doi: 10.1371/journal.pone.0003995 Vuilleumier, P., Henson, R. N., Driver, J., & Dolan, R. J. (2002). Multiple levels of visual object constancy revealed by event-related fMRI of repetition priming. Nature Neuroscience, 5(5), 491-499. Walther, D. B., Caddigan, E., Fei-Fei, L., & Beck, D. M. (2009). Natural scene categories revealed in distributed patterns of activity in the human brain. Journal of Neuroscience, 29(34), 10573- 10581. doi: 10.1523/JNEUROSCI.0559-09.2009 Weisberg, J., van Turennout, M., & Martin, A. (2007). A neural system for learning about object function. Cerebral Cortex, 17(3), 513-521. doi: 10.1093/cercor/bhj176 22
1609.01663
2
1609
2016-10-26T16:19:04
The canonical semantic network supports residual language function in chronic post-stroke aphasia
[ "q-bio.NC" ]
Current theories of language recovery after stroke are limited by a reliance on small studies. Here, we aimed to test predictions of current theory and resolve inconsistencies regarding right hemispheric contributions to long-term recovery. We first defined the canonical semantic network in 43 healthy controls. Then, in a group of 43 patients with chronic post-stroke aphasia, we tested whether activity in this network predicted performance on measures of semantic comprehension, naming, and fluency while controlling for lesion volume effects. Canonical network activation accounted for 22-33% of the variance in language test scores. Whole-brain analyses corroborated these findings, and revealed a core set of regions showing positive relationships to all language measures. We next evaluated the relationship between activation magnitudes in left and right hemispheric portions of the network, and how right hemispheric activation related to the extent of left hemispheric damage. Activation magnitudes in the each hemispheric network were strongly correlated, but four right frontal regions showed heightened activity in patients with large lesions. Activity in two of these regions (inferior frontal gyrus pars opercularis and supplementary motor area) was associated with better language abilities in patients with larger lesions, but poorer language abilities in patients with smaller lesions. Our results indicate that bilateral language networks support language processing after stroke, and that right hemispheric activations related to extensive left hemisphere damage occur outside of the canonical semantic network and differentially contribute to in their relationship to behavior depending on the extent of left hemispheric damage
q-bio.NC
q-bio
Title: The canonical semantic network supports residual language function in chronic post-stroke aphasia Authors: Joseph C. Griffis1, Rodolphe Nenert2, Jane B. Allendorfer2, Jennifer Vannest3, Scott Holland3, Aimee Dietz4, Jerzy P. Szaflarski2 Institutional Affiliations: University of Alabama at Birmingham Department of Psychology1, University of Alabama at Birmingham Department of Neurology2, Cincinnati Children's Hospital Medical Center, Cincinnati, OH3, University of Cincinnati Academic Health Center, Cincinnati, OH4 Corresponding Author Information: Joseph C. Griffis ([email protected]) Department of Neurology and UABEC, University of Alabama at Birmingham, 312 Civitan International Research Center, 1719 6th Avenue South, Birmingham, AL 35294- 0021 Abstract Current theories of language recovery after stroke are limited by a reliance on small studies. Here, we aimed to test predictions of current theory and resolve inconsistencies regarding right hemispheric contributions to long-term recovery. We first defined the canonical semantic network in 43 healthy controls. Then, in a group of 43 patients with chronic post-stroke aphasia, we tested whether activity in this network predicted performance on measures of semantic comprehension, naming, and fluency while controlling for lesion volume effects. Canonical network activation accounted for 22- 33% of the variance in language test scores. Whole-brain analyses corroborated these findings, and revealed a core set of regions showing positive relationships to all language measures. We next evaluated the relationship between activation magnitudes in left and right hemispheric portions of the network, and characterized how right hemispheric activation related to the extent of left hemispheric damage. Activation magnitudes in the each hemispheric network were strongly correlated, but four right frontal regions showed heightened activity in patients with large lesions. Activity in two of these regions (inferior frontal gyrus pars opercularis and supplementary motor area) was associated with better language abilities in patients with larger lesions, but poorer language abilities in patients with smaller lesions. Our results indicate that bilateral language networks support language processing after stroke, and that right hemispheric activations related to extensive left hemisphere damage occur outside of the canonical semantic network and differentially relate to behavior depending on the extent of left hemispheric damage. Keywords: fMRI, language recovery, stroke, lesion, right hemisphere Introduction Functional neuroimaging of language recovery after stroke Aphasia commonly occurs in patients with strokes affecting the left middle cerebral artery (LMCA) territory, and is one of the most debilitating consequences of stroke since language impairments affect nearly every aspect of daily life [Maas et al., 2012]. Language recovery after LMCA stroke is variable, with many survivors experiencing chronic deficits [Charidimou et al., 2014; Lazar et al., 2008; Pedersen, 1995]. Current theories of aphasia recovery are based primarily on evidence obtained from functional neuroimaging studies [Heiss and Thiel, 2006; Saur and Hartwigsen, 2012; Turkeltaub et al., 2011]. During early recovery, patients have been observed to show an up-regulation of right hemispheric responses to language tasks [Heiss et al., 1999; Saur et al., 2006] that is thought to reflect a transient compensatory mechanism triggered by the acute disruption of left hemispheric function by the ischemic event [Hartwigsen et al., 2013; Saur et al., 2006; Thiel et al., 2006]. In later stages of recovery, the gradual reinstatement of left hemispheric capacity for language processing is thought to decrease the need for right hemispheric compensation, resulting in a return of relatively balanced hemispheric activation patterns [Hamilton et al., 2011; Heiss and Thiel, 2006; Saur et al., 2006; Saur and Hartwigsen, 2012; Turkeltaub et al., 2011]. Thus, the general principle to be inferred from functional neuroimaging studies of language recovery after stroke might be summarized as follows: good long-term language outcomes depend primarily on the preservation and/or restoration of close-to-normal function in canonical brain language networks. While there is consistent support in the literature for a primary role of the canonical left hemispheric language networks in supporting language recovery after stroke, the role of the right hemisphere is less clear, particularly with regard to the chronic recovery phase. This is likely due, in part, to an oversimplified conceptualization of right hemispheric contributions to recovery in much of the literature [Turkeltaub et al., 2012]. Nonetheless, seemingly discrepant findings regarding the role of the right hemisphere pose an obstacle to developing a deeper understanding of how/when the right hemisphere contributes to language recovery. For example, there is evidence that activation of right inferior frontal [Belin et al., 1996; Griffis et al., 2016b; Rosen et al., 2000; Winhuisen et al., 2005; Winhuisen et al., 2007] and right superior temporal [Heiss et al., 1999; Karbe et al., 1998; Szaflarski et al., 2013] areas during language tasks are associated with poorer language function, although there is also evidence that activation of right inferior frontal [Mattioli et al., 2014; van Oers et al., 2010; Raboyeau et al., 2008; Saur et al., 2006] and right anterior superior temporal [Crinion and Price, 2005] regions may support residual language processing and contribute to language recovery. One explanation for such seemingly discrepant relationships is that when the left hemisphere is not sufficiently preserved, the initial reliance on right hemispheric language network homologues persists beyond early recovery, and the lack of left hemispheric involvement leads to poorer recovery relative to when left hemispheric function is successfully restored [Heiss et al., 1999; Heiss and Thiel, 2006; Karbe et al., 1998; Szaflarski et al., 2013; Turkeltaub et al., 2011]. In this case, the direction of the observed relationship between right hemispheric activation and language task performance might depend on the characteristics of the patient sample. For example, right hemispheric activation might show a positive relationship to language abilities in a sample that primarily consists of severely impaired patients with extensive left hemispheric damage, whereas a negative relationship might be observed in a sample that includes patients with various degrees of impairment and varying extents of left hemispheric damage [Saur and Hartwigsen, 2012]. Sample sizes in functional neuroimaging studies of aphasia recovery A limitation shared by many of the studies that form the foundation for current theories of recovery is that they typically utilize small patient samples [Saur and Hartwigsen, 2012], and this has the potential to (1) increase the risk of detecting spurious effects that do not generalize beyond the sample being studied, (2) reduce power to detect real effects, and (3) produce inflated estimates of the magnitudes of detected effects. Indeed, it has been previously noted in the context of neuroimaging that correlation estimates based on small samples are also unstable and can be disproportionately influenced by outlier data points [Poldrack, 2012]. Further, the power to detect between- subject effects is reduced in small samples [Yarkoni, 2009], and these relationships are often of primary interest in neuroimaging studies of language recovery after stroke. Even for real effects, small sample sizes inflate estimates of effect size, such that for a relationship with a population r of 0.3, the average observed sample r in samples of 20 patients is expected to be around 0.73 [Yarkoni, 2009]. To illustrate the prevalence of small sample sizes in functional neuroimaging studies (i.e. not including structural/lesion-mapping studies) of post-stroke aphasia, we performed a literature search using Google Scholar. Using the search terms "aphasia fmri", "aphasia pet", "aphasia neuroimaging", "aphasia recovery", "aphasia imaging", and "language recovery after stroke", we identified 84 functional neuroimaging (i.e. not including structural/lesion-mapping studies) studies published between 1995 and 2016 with accessible texts. Across all 84 studies, the average patient sample size was only about 10 patients (mean = 10.33, SD = 8.44; Figure 1), and 87% (73/84) had fewer than 20 patients. The single largest sample was reported by a study published this year [Geranmayeh et al., 2016] that collected data from 53 sub-acute patients (see Supplementary Material S1 for individual study sample sizes and a full reference list). While these estimates are based on a relatively coarse literature search, we note that they are in close agreement with the previous observation by Saur and Hartwigsen [2012] that neuroimaging studies of therapy-induced aphasia recovery typically feature typically less than 10 patients. Despite the limitations of small samples in individual studies, meta-analytic methods such as activation likelihood estimation (ALE) provide a means for identifying consistent effects in the literature [Wager et al., 2007]. However, we are aware of only one such published meta-analysis of the functional neuroimaging of aphasia recovery literature [Turkeltaub et al., 2011]. Larger studies, in conjunction with meta-analytic assessments of the published literature, are crucial for resolving discrepancies in the literature and for building generalizable theories of recovery that are based on robust empirical evidence. Statistical control for lesion volume in functional neuroimaging of aphasia recovery A second potential limitation shared by most functional neuroimaging studies of language recovery after stroke is a lack of statistical control for lesion volume effects on imaging-behavior relationships. This is somewhat surprising given that the use of statistical controls to reduce confounds related to lesion volume effects is relatively common in lesion-symptom mapping research [Rorden and Karnath, 2004; Schwartz et al., 2009; Zhang et al., 2014]. The logic behind using statistical controls to account for lesion volume effects is that since (1) larger lesions are often associated with more severe impairments [Allendorfer et al., 2012; Butler et al., 2014; Cheng et al., 2014; Karnath et al., 2004; Kümmerer et al., 2013; Meltzer et al., 2013; van Oers et al., 2010; Rorden and Karnath, 2004; Szaflarski et al., 2013; Yarnell et al., 1976], and (2) larger lesions have a higher probability of including both task-relevant and task-irrelevant voxels [Karnath et al., 2004; Rorden and Karnath, 2004; Schwartz et al., 2009; Zhang et al., 2014], lesion volume effects have the potential to introduce bias into measurements of lesion-behavior relationships at voxels that are primarily damaged in patients with large lesions [Karnath et al., 2004; Zhang et al., 2014]. In functional neuroimaging, the probability that a voxel is active during task performance depends on the probability that the tissue at that voxel is spared, and the probability that the tissue at a given voxel in the left hemisphere is spared depends on the size of the lesion. Therefore, the potential for lesion volume confounds is logically extendable to relationships between functional neuroimaging activation and behavior, and it is possible that a lack of controls for such effects could introduce bias into functional neuroimaging studies of recovery. Study aims and hypotheses In the current study, we first tested what we consider to be the primary prediction of current models of language recovery after stroke using a relatively large (n=43) sample of patients with chronic (> 1 year) post-stroke aphasia and a sample (n=43) of demographically matched healthy controls. Specifically, we tested the prediction that long-term language outcomes depend on the preservation/restoration of language task- driven activation in canonical language networks. To test this prediction, we first defined the canonical semantic network (CSN) as the set of regions that were more strongly activated by semantic decisions relative to tone decisions in the healthy controls. This bilateral but predominantly left-lateralized network is well-suited for testing the prediction that residual language abilities depend on the preservation/recovery of activity within distributed canonical language networks as it is (1) well-characterized in healthy individuals, and (2) consists of distributed brain regions that include the left angular gyrus (AG)/posterior inferior parietal lobule (pIPL), left middle temporal gyrus (MTG), left parahippocampal gyrus (PHG), left dorsal superior frontal gyrus (dSFG), left inferior frontal gyrus (IFG), and left posterior cingulate cortex (PCC) [Binder et al., 1997; Binder et al., 1999; Binder et al., 2008; Binder et al., 2009; Kim et al., 2011; Szaflarski et al., 2002; Xu et al., 2016]. Thus, we tested whether task-evoked activity in the CSN identified in controls predicted in-scanner performance on the semantic decision task, out-of-scanner performance on verbal fluency tasks, and out-of-scanner performance on a picture naming task in the patient group (with and without controls for lesion volume effects). We note that while the in-scanner and out-of-scanner tasks differ in their emphases on speech comprehension and speech production, respectively, previous studies suggest that activity evoked by speech comprehension tasks may also relate performance on other measures of language function that include speech production tasks [e.g. Saur et al., 2006; Szaflarski et al., 2013]. Other previous studies also suggest that speech production networks (i.e. regions activated during word/verb generation) are also activated (albeit somewhat less strongly) during sentence comprehension tasks [Piervincenzi et al., 2016], and that estimates of language lateralization based on activations measured during both speech production and auditory semantic decision tasks show similar relationships to estimates based on intracarotid amobarbital measurements [Szaflarski et al., 2008]. Additionally, we tested the explanation that strong right hemispheric activation in chronic patients reflects compensation driven by a reduced capacity of the left hemisphere to perform language-related processing. As noted in the previous paragraph, the CSN has bilateral components [e.g. Binder et al., 2009], and thus activation of the right hemispheric network might be expected to increase to compensate for reduced left hemispheric network function [e.g. Hartwigsen et al., 2013]. While previous studies have investigated how activity in the right hemisphere relates to the effects of damage to specific left hemispheric regions [Blank et al., 2003; Sims et al., 2016; Turkeltaub et al., 2011], we are not aware of any studies that have directly measured the how activation in the right hemisphere relates to the extent of damage sustained by the left hemisphere. Further, the notion that language-task evoked activation in the right hemisphere is most pronounced in patients with extensive left hemispheric lesions is commonly discussed in the literature [Hamilton et al., 2011; Heiss and Thiel, 2006]. Despite its intuitive appeal, there appears to be little empirical evidence to support this claim. For example, a recent review of the mechanisms of language recovery after stroke [Hamilton et al., 2011] cites only two reports to support the unambiguous assertion that "larger lesions involving eloquent cortex of the left hemisphere are associated with greater recruitment of the right hemisphere during language tasks". One is an earlier review [Heiss and Thiel, 2006] that discusses this idea within the context of a larger hierarchical model of recovery but does not itself provide empirical support for the veracity of this claim in stroke patients, and the other [Kertesz et al., 1979] is a computerized tomography study that shows more extensive left hemispheric damage to be predictive of more severe aphasia and poorer language recovery after stroke, but that does not relate lesion size to any measurement of functional activation. Therefore, we sought to empirically test the hypotheses that right hemispheric language task activation in chronic stroke patients reflects compensation driven by reduced left hemispheric function, and that this effect is most pronounced in patients with the most extensive left hemispheric damage. We expected, based on the literature discussed above, that higher levels of right CSN activation during semantic decisions would be associated with (1) lower levels of left CSN activation, and (2) more extensive left hemispheric damage. As noted above, the CSN is predominantly left-lateralized (i.e. the right and left CSN are not mirror- symmetric), and so we also assessed these effects in right hemispheric regions mirroring the left CSN. Finally, because we are not aware of any studies that have directly measured the effect of left hemispheric lesion volume on regional activation in the right hemisphere, we characterized regions in the right hemisphere where task-driven responses correlated with left hemispheric lesion volume. To address the question of whether activity in these regions might reflect beneficial compensation that is most pronounced for patients with extensive left hemispheric damage, we further assessed whether relationships between activity in these regions and language measures depended on the extent of left hemispheric damage. All study procedures were approved by the Institutional Review Boards of the Materials and Methods Participants participating institutions and were performed in accordance with Declaration of Helsinki ethics principles and principles of informed consent. For the current study, we used functional MRI data collected from 43 chronic post-stroke aphasia patients and 43 healthy controls. Imaging and behavioral data for the post-stroke aphasia patients were collected as part of several separate studies performed by our laboratory. Prior to inclusion, all participants were screened to exclude individuals that had diagnoses of degenerative/metabolic disorders, had severe depression or other psychiatric disorders, were pregnant, were not fluent in English, or had any contraindication to MRI/fMRI. Patients were included in the current study if they had a single left hemispheric stroke resulting in aphasia at least 1 year prior to data collection. Data for the control participants were selected from a database of 150 healthy individuals collected as part of several studies performed by our laboratory, and were selected based on age group (19- 29, 30-39, 40-49, 50-59, 60+), handedness as determined by the Edinburgh Handedness Inventory [Oldfield, 1971], and sex to minimize differences in demographics with respect to the patient group. Participant demographics are shown in Table 1. A more detailed characterization of patient demographics is provided in the Supplementary Material (S2). Language measures Prior to MRI scanning, all participants were administered a battery of neuropsychological language assessments. All participants performed the Boston Naming Test (BNT) [Kaplan et al., 2001], Semantic Fluency Test (SFT) [Kozora and Cullum, 1995], and Controlled Oral Word Association Test (COWAT) [Lezak et al., 1995]. The BNT requires patients to name a series of black and white line drawings that contain both animate and inanimate items that vary in frequency of use (e.g. bed vs. abacus), and the number of correctly named pictures serves as a measure of naming ability. The SFT requires patients to generate as many words as they can think of that fit a given category prompt (animals/fruits and vegetables/things that are hot) within a one-minute time limit, and the number of words generated serves as a measure of category fluency. The COWAT requires patients to generate as many words as they can think of that begin with a particular letter (C/F/L) within a one-minute time limit, and the number of words generated serves as a measure of phonemic fluency. Performance scores for the COWAT and SFT were very strongly correlated across patients (r=0.92), so they were averaged together to define a single combined measure of verbal fluency. Performance scores for the verbal fluency measure were correlated, albeit less strongly, with performance scores for the BNT (r=0.76). Individual patient language task data are provided in the Supplementary Material (S2). Neuroimaging data collection MRI data were collected at the University of Alabama at Birmingham using a 3T head-only Siemens Magnetom Allegra scanner located in the Civitan International Research Center Functional Imaging Laboratory. These data consisted of 3D high- resolution T1-weighted anatomical scans (TR/TE = 2.3 s/2.17 ms, FOV = 25.6×25.6×19.2 cm, matrix = 256x256, flip angle = 9 degrees, slice thickness = 1mm), and two T2*-weighted gradient-echo EPI pulse functional scans (TR/TE = 2.0 s/38.0 ms, FOV = 24.0x13.6x24.0, matrix = 64x64, flip angle = 70 degrees, slice thickness = 4 mm, 165 volumes per scan). MRI data were also collected at the Cincinnati Children's Hospital Medical Center on a 3T research-dedicated Phillips MRI system located in the Imaging Research Center. These data consisted of 3D high-resolution T1-weighted anatomical scans (TR/TE = 8.1 s/2.17 ms, FOV = 25.0×21.0×18.0 cm, matrix = 252x211, flip angle = 8 degrees, slice thickness = 1mm) and two T2*-weighted gradient-echo EPI pulse sequence functional scans (TR/TE = 2.0 s/38.0 ms, FOV = 24.0x13.6x24.0, matrix = 64x64, flip angle = 70 degrees, slice thickness = 4 mm, 165 volumes per scan). Functional MRI scans were acquired while participants completed alternating 30 second blocks of a semantic decision/tone decision task. This fMRI task was selected because it has been previously shown to result in robust activation in canonical areas involved in semantic language processing [Binder et al., 1997; Kim et al., 2011], and has been used extensively to evaluate language network activation in healthy and diseased populations that include patients with chronic post-stroke aphasia [e.g. Eaton et al., 2008; Szaflarski et al., 2008; Szaflarski et al., 2011]. The active condition (semantic decision -- SD) was performed 5 times during each scan. Each block of the active condition consisted of 8 trials where participants were presented with spoken English animal names. On each trial, participants decided if the presented animal met the criteria "native to the United States" and "commonly used by humans". If both criteria were satisfied, then the participants responded "1" by using their non-dominant hand to press a button. If both criteria were not satisfied, then the participants responded "2" by using their non-dominant hand to press a second button. The control condition was performed 6 times during each scan. Each block of the control condition (tone decision – TD) consisted of 8 trials where participants were presented with brief sequences of four to seven 500- and 750-Hz tones. On each trial, participants decided if the sequence contained two 750-Hz tones. If the sequence contained two 750 Hz tones, then they pressed the button designated "1" with their non-dominant hand. Otherwise, they pressed the button designated "2" with their non-dominant hand. Each scan lasted 7 minutes and 15 seconds. Prior to completing the task in the MRI scanner, all participants confirmed their understanding of the task by performing a mock run that included a sequence of five sets of tones followed by a sequence of five nouns designating different animals. In-scanner task data were not collected for 4 patients due to hardware issues, and they were excluded from analyses investigating relationships between fMRI activation and in-scanner performance. Neuroimaging data processing All MRI data were processed using Statistical Parametric Mapping (SPM) [Friston et al., 1995] version 12 running in MATLAB r2014b (The MathWorks, Natick MA, USA). Functional MRI scans were pre-processed using a standard pre-processing pipeline consisting of slice-time correction, realignment/reslicing, co-registration of the fMRI data to the corresponding anatomical scan, unified segmentation with optimized tissue priors for lesioned brains and normalization of the anatomical scan to MNI space [Ripollés et al., 2012; Seghier et al., 2008], normalization of the functional scan to MNI space using the transformation applied to the anatomical scan, and spatial smoothing using an 8mm full-width half maximum Gaussian kernel. This pipeline enables accurate template registration and normalization even for patients with structural abnormalities such as those observed in stroke patients [Ripollés et al., 2012]. To reduce the potential for motion-related artifacts, functional scans were motion-corrected by replacing volumes with >0.5mm motion with an interpolated volume created from adjacent volumes [Mazaika et al., 2005]. Lesion probability maps were created for each patient using a voxel-based naïve Bayes classification algorithm that was developed by our laboratory and is implemented in the lesion_gnb toolbox for SPM12 [Griffis et al., 2016a]. While automated classification with this method compares favorably to manual lesion delineation for large and small lesions [Griffis et al., 2016a], we opted to manually threshold the resulting posterior probability maps to ensure that they precisely reflected the extent of the lesion. Our decision to manually threshold the resulting posterior probability maps was primarily motivated by the potential for automated methods to introduce false positive voxel clusters in patients with small lesions [e.g. Griffis et al., 2016a; Wilke et al., 2011], and while an arbitrary cluster threshold (i.e. 100 voxels) is often sufficient to remove these clusters, it is not guaranteed. Further, as we note in our report describing the validation of our method [Griffis et al., 2016a] it is important to inspect the lesion masks as a quality control step, and by manually thresholding the probability maps produced by the automated classification procedure, we were able to ensure that the final lesion masks accurately reflected each patient's lesion. The resulting binary lesion masks were used to estimate lesion volume and lesion-ROI overlaps, and were used in all further lesion analyses. Lesion frequencies across all 43 patients are shown in Figure 2A. Individual patient lesion images are provided in the Supplementary Material (S2). For each participant, the pre-processed fMRI data were fit to a general linear model (GLM) where task blocks were modeled as boxcar regressors convolved with a canonical hemodynamic response function (HRF). In order to account for variability in the time-to-peak of the HRF, time and dispersion derivatives were included as basis functions [Meinzer et al., 2013]. Note that since data were collected using a blocked design, separate modeling of correct vs. incorrect trials was not possible. Linear contrast estimate maps were then computed to quantify the difference in activation magnitudes between the active and control conditions. These contrast maps were used for all further functional MRI analyses. Statistical Analyses Group differences on behavioral measures were assessed for descriptive purposes using two-tailed independent samples t-tests with degrees of freedom adjustment for unequal variances. To identify regions that showed significant group-level activation for each group, the first-level contrast estimate maps from all individuals in each group were entered into separate second-level t contrasts quantifying the difference in peak HRF magnitude between the SD and TD conditions. Regional activation was considered significant if it survived a combined voxel-level intensity (p-value) threshold of 0.01 and a cluster-level extent threshold of p<0.05 corrected (kcrit=99 voxels) to control the whole-brain family- wise error rate (FWE) at 0.05 as determined by 1000 Monte Carlo simulations [Slotnick et al., 2003]. The thresholded map obtained for the control group was then binarized, creating a CSN region-of-interest (ROI) mask where statistically significant voxels had a value of 1 and all other voxels had a value of 0. For each patient, the mean contrast estimate across all voxels within this CSN ROI was then extracted from their first-level contrast estimate map, quantifying the average magnitude of the task-driven response in the CSN. These estimates were used in subsequent ROI correlation analyses. Group differences in activation within the CSN were assessed for descriptive purposes using an independent samples t-test with degrees of freedom adjustment for unequal variances. Lesion-ROI overlaps were calculated for each patient using the lesion masks created during pre-processing. The correlation between total lesion volumes and lesion-ROI overlaps indicated that lesion-ROI overlap was a close linear transformation of lesion volume (r=0.93). To assess the unique relationships between activity in the CSN and language function in chronic patients, the mean contrast estimate quantifying activation within the HC-defined CSN was then entered as an independent variable in 3 separate partial correlation analyses. The dependent variables for each analysis were the performance scores for the in-scanner SD task (% correct responses) and performance scores for the out-of-scanner naming and fluency tasks. Each model included total lesion volume as a covariate to account for variance attributable to lesion volume effects. Analyses were repeated without lesion volume control. Linear correlation analyses assessing imaging- behavior relationships were also performed for the control group using the CSN ROI. To fully characterize relationships between regional language task activation and performance on the in-scanner and out-of-scanner language tasks, three whole-brain multiple regressions (one for each language task) were performed that each included lesion volume as a covariate. Because patient data were collected on different scanners, all correlational analyses for the patient group were repeated with scanner included as a covariate, with and without lesion volume control. These analyses are provided in the Supplementary Material (Supplementary Figure 4). To test whether the magnitude of activation in right hemispheric portions of the CSN was increased in patients with lower levels of activation in left hemispheric portions of the network, the CSN ROI mask was split into left and right hemispheric components. For each patient, the mean contrast estimate was then extracted from each hemispheric ROI as described above. Linear correlations were assessed between the mean contrast estimates obtained from each hemisphere with and without lesion volume control. Correlations were also assessed between left hemispheric and right hemispheric activation in controls. To explicitly test for this effect in right hemispheric homologues of the left hemisphere network and control for differences in extent between left and right hemispheric portions of the CSN, these analyses were also repeated using a right hemispheric ROI that was defined by simply mirroring the left-hemispheric ROI to the right hemisphere. To assess whether out-of-network right hemispheric homologues of the left hemispheric CSN (i.e. regions that were activated in the left, but not right hemisphere in controls) differed from the right hemispheric CSN with regard to their relationship to left hemispheric CSN activation, a right hemispheric out-of-network homologue ROI was created by subtracting the right hemispheric CSN from the mirrored left hemispheric CSN (see Figure 4A for illustrations of each ROI). Differences in activation between left and right hemispheric networks were assessed with dependent samples t-tests for each group. Between-group differences in activation for left and right hemispheric networks (and differences in activation between left and right hemispheric networks) were assessed with unequal variance t-tests. Between-group differences in the strengths of correlations between left and right CSN ROIs were assessed using z-tests [Fisher, 1921]. Lastly, to assess the effect of lesion size on task-driven activation in the right hemisphere, we performed an additional whole-brain linear regression analysis that was restricted to the right hemisphere and included only total lesion volume as a predictor. Post-hoc moderation analyses were performed for right hemisphere clusters that showed significant positive correlations between activation magnitudes and lesion volume in order to assess whether activity in these regions reflected beneficial compensation in patients with larger lesions. For each identified cluster, three separate multiple regression analyses (one for each language measure) were performed that each included mean activation magnitudes for that cluster, total lesion volume (entered as a proportion of the maximum lesion volume), and the interaction term between mean cluster activation and lesion volume. For these post-hoc analyses, only the interaction effects were of interest, since we our hypothesis was that the benefit of recruiting these regions would depend on the extent of left hemispheric damage. Statistical tests were two-tailed and significance thresholds were set as follows. Results for group comparisons and for each set of ROI-based analyses were considered significant if they survived a Bonferroni-Holm step-down correction procedure to control the FWE at 0.05 across each set of tests, since this method provides good protection against Type I errors and better power than standard Bonferroni correction [Aickin and Gensler, 1996]. T-contrast maps for the effects of interest from each of the whole-brain multiple regression analysis were thresholded at p<0.01, uncorrected at the voxel level and multiple comparisons corrected at the cluster level to control the cluster-wise FWE at 0.01 (kcrit = 126 voxels, as determined by 1000 Monte Carlo simulations) for each contrast. Since power is often lower for analyses of moderation [Jaccard et al., 1990], the post-hoc moderation analyses were multiple comparisons corrected using a more lenient False Discovery Rate (FDR) correction threshold of 0.10 [Benjamini and Hochberg, 1995] computed across the interaction term p-values for all moderation models. We note that a recent report [Eklund et al., 2016] has called into question the validity of cluster-based correction methods in fMRI research, particularly when they are used in conjunction with voxel-level thresholds above 0.001. Our choice to use the voxel- level threshold of 0.01 was primarily motivated by its prevalence in the neuroimaging literature of post-stroke aphasia, and because detection power is reduced for group analyses of highly heterogeneous groups such as stroke patients [e.g. Meinzer et al., 2013]. In addition, we note that for the whole-brain regression analyses, we employed a more stringent whole-brain cluster-correction threshold of 0.01 (as opposed to the conventional 0.05 threshold), even when our analyses were restricted to only the right hemisphere. Thus, we have attempted to balance providing control over false positives while maintaining sufficient power to detect effects of interest. In addition, we provide voxel-wise FDR thresholded activation maps and cluster/peak statistics for each group in Supplementary Figure 4 to (1) demonstrate that similar activation patterns are detected at stringent voxel-wise correction thresholds, and (2) provide more precise anatomical information about peak activation locations for each group. Results Comparisons of overall CSN activation and task performance between patients and controls Patients performed more poorly on language tasks than controls (in-scanner SD % correct: t74.77=-4.71, p<0.001, corrected; BNT: t43.38=-6.80, p<0.001, corrected; fluency: t82.28=-13.44, p<0.001, corrected). Summary statistics are shown in Figure 2B. The semantic networks recruited by each group are shown in Figure 2C. Cluster peak co- ordinates and statistics for regions activated by each group are provided in Table 2. Controls showed significantly higher levels of overall activation in the CSN than patients (t67.72 = 3.98, p<0.001, corrected). Region of interest behavioral correlations Total lesion volume showed negative relationships with each of the language measures (in-scanner SD correct: r=-0.22, p=0.19, corrected; fluency: r =-0.58, p<0.001, corrected; naming: r=-0.49, p=0.004, corrected), although the relationship was not significant for the in-scanner SD task. Lesion volume showed a trend-level but non- significant negative relationship with mean activation in the CSN (r=-0.27, p=0.15, corrected). The partial correlation analyses revealed significant positive relationships between activity within the CSN and in-scanner performance on the SD task (partial r=0.47, p=0.008, corrected), out-of-scanner performance on the fluency measure (partial r=0.62, p<0.001, corrected), and out-of-scanner performance on the naming measure (partial r=0.49, p=0.005, corrected) that were linearly independent of the effects of total lesion volume (Figure 2D). Relationships between CSN activation and in-scanner performance on the SD task (r=0.51, p=0.005, corrected), out-of-scanner performance on the fluency measure (r=0.65, p<0.001, corrected), and out-of-scanner performance on the naming measure (r=0.55, p=0.001, corrected) were stronger without lesion volume control. These results are shown in Figure 2D. Additional analyses assessed relationships between mean CSN activation magnitudes and behavioral measures in controls. Mean CSN activation magnitudes showed a significant positive relationship to performance on the in-scanner SD task (r=0.37, p=0.03, corrected), but did not show significant relationships with performance on the fluency (r=0.09, p=1.0, corrected) or naming (r=0.09, p=1.0, corrected) tasks (Figure 2D). Whole-brain behavioral correlations The results for the whole-brain multiple regressions for each language measure that included total lesion volume as a covariate are shown in Figure 3A. Cluster peak locations and statistics for each whole-brain regression are provided in Table 3. Figure 3B shows regions that showed positive relationships between activation and performance on all three language measures, and Figure 3C shows regions that showed positive relationships between activation and any of the language measures relative to the locations of regions associated with the CSN identified in healthy controls. Whole-brain analyses repeated without including lesion volume as a covariate showed very similar results (Supplemental Tables 7-9). Correlations between left and right hemispheric activation magnitudes Partial correlation analyses (controlling for lesion volume) for the patient group revealed a strong positive relationship between activation magnitudes in the left CSN and activation magnitudes in the right CSN (partial r=0.83, p<0.001, corrected). Nearly identical results were obtained without lesion volume control (r=0.83, p<0.001, corrected). A strong positive relationship was also observed between activation magnitudes in the left CSN and activation magnitudes in the mirrored left CSN in the right hemisphere with (partial r=0.81, p<0.001, corrected) and without (r=0.80, p<0.001, corrected) lesion volume control. However, correlations were much weaker between activation magnitudes in the left CSN and activation magnitudes in out-of-network homologues of the left hemispheric CSN in the right hemisphere with (partial r= 0.66, p<0.001, corrected) and without (r=0.65, p<0.001, corrected) lesion volume control. Left and right hemispheric ROIs are shown in Figure 4A. Scatterplots illustrating the relationships between residual left and right hemispheric activation magnitudes (i.e. after partialling out lesion volume) are shown in Figure 4B (top row). Correlation analyses for the control group also revealed a strong positive relationship between activation magnitudes in the left CSN and activation magnitudes in the right CSN (r=0.75, p<0.001, corrected). A weaker relationship was observed between activation magnitudes in the left CSN and activation magnitudes in the mirrored left CSN in the right hemisphere (r=0.43, p=0.008, corrected). No significant relationship was observed between activation magnitudes in the left CSN and activation magnitudes in out-of-network homologues of the left CSN in the right hemisphere (r=-0.04, p=0.79, corrected). Scatterplots illustrating these relationships are shown in Figure 4B. The strengths of correlations between left CSN and right CSN activation magnitudes did not significantly differ between patients and controls (z=0.52, p=0.6, corrected), but the strengths of correlations between activation magnitudes in the left CSN and mirrored left CSN in the right hemisphere (z=2.98, p=0.006, corrected) and between activation magnitudes in the left CSN and out-of-network homologues of the left CSN in the right hemisphere (z=3.72, p<0.001, corrected) were significantly stronger in patients than in controls. Comparisons of left and right hemispheric activation magnitudes Activation magnitudes for each group are shown in Figure 4C. Neither patients (t42=0.96, p=0.34, corrected) nor controls (t42=1.72, p=0.19, corrected) showed significant differences in activation magnitudes between the left CSN and right CSN. Both patients (t42=3.89, p=0.001, corrected) and controls (t42=9.22, p<0.001, corrected) showed significantly larger activation magnitudes in the left CSN compared to the mirrored left CSN in the right hemisphere. Both patients (t42=5.42, p<0.001, corrected) and controls (t42=12.92, p<0.001, corrected) also showed significantly larger activation magnitudes in the right CSN compared out-of-network homologues of the left CSN in the right hemisphere. Patients had significantly lower activation magnitudes in both the left (t66.37=- 3.60, p<0.001, corrected) and right (t74.28=-4.23, p<0.001, corrected) CSN than controls. Patients also had significantly lower activation magnitudes in the mirrored left CSN in the right hemisphere (t67.17=-2.92, p=0.01, corrected). Patient and control activation magnitudes in out-of-network homologues of the left CSN in the right hemisphere did not significantly differ (t60.15=0.47, p=1.0, corrected). The difference between activation in the left CSN and the right CSN was not significantly larger in controls than in patients (t81.89=0.38, p=1.0, corrected). Controls showed a significantly larger difference in activation between the left hemispheric CSN and the mirrored left CSN in the right hemisphere (t83.82=3.99, p<0.001, corrected). Controls also showed a significantly larger difference between activation magnitudes in the right CSN and out-of-network homologues of the left CSN in the right hemisphere (t82.15=6.26, p<0.001, corrected). Lesion volume effects on activation magnitudes Activation magnitudes in both the left CSN (r=-0.24, p=0.35, corrected) and right CSN (r=-0.31, p=0.18, corrected) showed non-significant negative relationships to lesion volume. A weaker negative relationship was observed between lesion volume and activation magnitudes in the mirrored left CSN in the right hemisphere (r=-0.12, p=0.90, corrected). Lesion volume effects showed a near-zero correlation with activation magnitudes in out-of-network homologues of the left CSN in the right hemisphere (r=- 0.01, p=0.97, corrected). These results are shown in Figure 4D. The results from the whole-brain regression analysis assessing the effects of lesion volume on activation in the right hemisphere are shown in Figure 5A. Cluster statistics and peak locations are displayed in Table 4. All of the regions showing positive relationships with lesion volume fell outside of the canonical semantic network in right frontal cortex (Figure 5B). Notably, with the exception of the right middle frontal gyrus (MFG), these regions did not correspond to out-of-network homologues of the left CSN. In contrast, several of the regions showing negative relationships between activation and lesion volume overlapped with the CSN (Figure 5B). To assess whether there was a benefit of activating out-of-network right- hemisphere regions that depended on the extent of left hemispheric damage, we performed additional post-hoc regression analyses to assess whether the lesion size might moderate the effect of activity in each right hemispheric cluster showing positive correlations to total lesion volume (hot clusters in Figure 5A). This revealed significant interactions between lesion volume and right SMA activation for the fluency (t39=2.97, FDRp = 0.04) and naming (t39=2.82, FDRp = 0.04) measures (Figure 5C), and between lesion volume and right IFG pars opercularis activation for the fluency (t39=2.50, FDRp = 0.07) measure. For each of the observed interaction effects, higher levels of activation in the right hemispheric cluster was associated with lower language test scores for patients with smaller lesions, but with higher language test scores for patients with larger lesions (Figure 5C). Note that the high and low values shown in Figure 5C are calculated for +1 and -1 standard deviations from the mean of each variable. Discussion Functional neuroimaging studies of post-stroke aphasia are foundational to theories of language recovery after stroke, but this theoretical foundation is primarily composed of small studies. To address this shortcoming and assess the validity of predictions based on current theories of recovery, we tested what we consider to be their primary prediction – that successful long-term language recovery depends on the preservation and/or restoration of language processing in canonical language networks. Importantly, we tested this prediction in one of the largest samples of post-stroke aphasia patients studied to date in the functional neuroimaging literature, and utilized statistical controls to reduce the potential for confounding effects related to lesion volume on imaging-behavior relationships. We also attempted to bring clarity to contested issues in the literature – how language task-driven activity in the unaffected right hemisphere relates to left hemispheric damage and function, and whether the recruitment of right hemispheric regions by patients with extensive left hemispheric damage supports residual language function. We discuss our results and their relation to the broader literature in the following sections. Canonical language network contributions to post-stroke language function Drawing from the broader functional neuroimaging literature of language recovery after stroke, we expected that activation of the CSN for language processing would positively predict language functions in chronic patients. The results of both our ROI-based and whole-brain analyses matched this expectation, and indicate that activation of the CSN is a moderate-to-strong predictor of language function in chronic patients (Figures 2 and 3). While activation magnitudes in this network only predicted performance on the in-scanner task for controls, they predicted performance on all three language measures for patients (Figure 2D). Further, our whole-brain analyses of the patient data revealed that the positive relationships between regional task-driven activation magnitudes and language task performance were primarily localized within or adjacent to the CSN identified in controls (Figure 3). Thus, our results corroborate the general implications of previous functional neuroimaging studies of language outcomes in patients with LMCA stroke [Fridriksson et al., 2010; Fridriksson et al., 2012; Heiss et al., 1999; Karbe et al., 1998; van Oers et al., 2010; Rosen et al., 2000; Saur et al., 2006; Szaflarski et al., 2013]. Importantly, this result supports the emphasis of current theory on the preservation and/or restoration of function in canonical language networks as a key factor that enables successful long-term language recovery after stroke. Notably, our results indicate that despite the fact that the auditory semantic decision task lacks an expressive language component, the level of CSN activation evoked was a good predictor of expressive language capacity in our patient sample (Figure 2D/Figure 3). As noted in the introduction, previous studies suggest that relationships between activation evoked by auditory (i.e. receptive) language tasks and language function in chronic stroke patients may not be specific to receptive language functions [e.g. Saur et al., 2006; Szaflarski et al., 2013]. Our results suggest that activation in the CSN evoked by auditory semantic decisions may provide an index of general language network preservation/recovery. This interpretation is consistent with evidence that the abrupt and catastrophic disruption of neural function by stroke leads to widespread disruptions of communication and regulation in distributed brain networks [Baldassarre et al., 2016; Geranmayeh et al., 2016; He et al., 2007; Ovadia-Caro et al., 2013; Siegel et al., 2016], and that the preservation/restoration of typical function in canonical language networks is indicative of successful language recovery [Heiss et al., 1999; Saur et al., 2006]. Indeed, post-stroke deficits in complex cognitive functions [Siegel et al., 2016] that include language [Geranmayeh et al., 2016] and attention [Baldassarre et al., 2016; He et al., 2007] may be conceptualized as behavioral manifestations of dysfunction in large-scale functional brain networks. Along these lines, optimal functional recovery after stroke may depend on the preservation/restoration of functional dynamics that most strongly resembles those observed in the pre-stroke brain [Carter et al., 2012]. Thus, the magnitude of task-evoked responses in the distributed CSN may index the degree to which pre-stroke functional dynamics are preserved/restored in patients with chronic post-stroke aphasia. While discussions of how distributed networks contribute to recovery from stroke often focus on measures of intrinsic (i.e. resting state) network function [e.g. Carter et al., 2012], there is substantial evidence that task-evoked brain networks are strongly influenced by the intrinsic network architecture [Binder et al., 1999; Binder, 2012; Cole et al., 2014; Dosenbach et al., 2006; Dosenbach et al., 2007; Fox et al., 2005; Fox et al., 2006; Muhle-Karbe et al., 2015; Xu et al., 2016], and thus it is possible that the presence of strong activation in the CSN during auditory semantic decisions may be indicative of well-preserved and/or successfully restored intrinsic network dynamics. We must emphasize, however, that this explanation is tentative and must be confirmed by future studies that are capable of relating task-evoked responses in canonical language networks to intrinsic network dynamics in patients with chronic post- stroke aphasia. Co-activation of left and right hemispheric networks and right hemispheric effects of lesion volume Based on current models of language recovery after stroke [Hamilton et al., 2011; Heiss and Thiel, 2006; Saur et al., 2006; Saur and Hartwigsen, 2012], we expected that patients with lower levels of activation in the left CSN would show higher levels of activation in the right CSN. Assuming that this would reflect compensatory up-regulation of the right CSN following extensive left hemispheric damage, we also expected that activation in the right CSN would be increased in patients with larger left hemispheric lesion volumes. However, our results did not bear out this prediction. Rather, we found that for both patients and controls, activation magnitudes in the right CSN were strongly positively correlated with activation magnitudes in the left CSN (Figure 4B). Further, in patients, activation magnitudes in both the left and right CSNs showed negative but non- significant correlations with lesion volume (Figure 4C). This suggests that the right CSN and left CSN form a coherent functional unit that is activated by semantic processing. Indeed, a basic role of certain right hemispheric regions in language processing is suggested by findings that right hemispheric activation during semantic decision [Donnelly et al., 2011], semantic comprehension [van Ettinger-Veenstra et al., 2010], sentence completion [van Ettinger-Veenstra et al., 2012], and word fluency tasks [van Ettinger-Veenstra et al., 2012] correlates with various measures of language function in healthy individuals. Notably, a recent ALE meta-analysis of 12 functional neuroimaging studies of aphasic patients (total n=105) and healthy controls (total n=129) suggests that specific regions in both hemispheres are consistently recruited by aphasic patients across different language task paradigms [Turkeltaub et al., 2011]. Turkeltaub and colleagues [2011] thus proposed that activity in most right hemispheric regions, with the notable exception of the IFG pars triangularis, likely reflects either compensatory recruitment or co-activation with homotopic areas in the left hemisphere. Our finding that activation magnitudes in the left and right CSN were strongly correlated suggests that at least with regard to the CSN, activity in right hemispheric portions of canonical networks likely reflects co- activation rather than compensation for left hemispheric damage. The finding that neither patients nor controls showed significant differences between the left and right CSN supports this conclusion. However, we also found that both patients and controls showed significantly higher activation in the left CSN than in homologous areas in the right hemisphere, and this effect was most pronounced for homologous areas that were not associated with the right CSN (Figure 5C). Further, both patients and controls showed significantly higher levels of activity in the right CSN compared to these out-of-network left CSN homologues (Figure 5C). This suggests that co-activations are spatially constrained to the subset of right hemispheric regions that typically activate as part of the CSN. While the correlation between activation magnitudes in the left CSN and out-of-network right hemispheric homologues was substantially weaker than the correlation between activation magnitudes in the left CSN and right CSN in patients (Figure 5B), controls showed essentially no relationship between activation magnitudes in the left CSN and out-of-network right hemispheric homologues (Figure 5B). This suggests that inter- hemispheric co-activations may be less spatially constrained in stroke patients than in healthy controls. Speculatively, this may reflect weakened inter-hemispheric inhibition, as this has been proposed as a potential source of increased activations in right hemispheric homologues of left hemispheric language areas in patients with post-stroke aphasia [Hamilton et al., 2011; Heiss and Thiel, 2006]. Nonetheless, our data support a functional distinction between right hemispheric regions that show robust co-activation with distributed language networks and out-of- network right hemispheric homologues of left hemispheric language areas, although conclusions regarding the source of this distinction cannot be drawn from this study. Future studies using task-based (i.e. effective) connectivity metrics to assess how activation in these regions relate to interactions between left and right hemispheric networks are necessary to address such questions. Right hemispheric compensation in patients with extensive left hemispheric damage Recent evidence from structural MRI studies also suggests a compensatory role of certain right hemispheric regions in supporting language function in chronic patients. For example, grey matter volume in right dorsal stream temporo-parietal areas is increased in chronic post-stroke aphasia patients relative to both healthy controls and chronic left hemispheric stroke patients without aphasia [Xing et al., 2015]. Importantly, higher grey matter volume in these regions is associated with better residual language functions in chronic post-stroke aphasia patients [Xing et al., 2015]. Similarly, increased fractional anisotropy (FA) has been observed in the right IFG pars opercularis for chronic post- stroke aphasia patients relative to healthy controls, and higher FA in this region is associated with better speech production abilities for patients [Pani et al., 2016]. Our findings add to this literature by showing that functional activation in specific right frontal regions during language task performance is associated with better expressive language abilities in chronic post-stroke aphasia patients with extensive left hemispheric damage. Notably, larger lesions were associated with higher levels of activity in several right frontal areas, including the right inferior frontal gyrus pars opercularis and supplementary motor area, both of which were not part of the CSN (Figure 5A/B). Activation magnitudes in the right IFG pars opercularis and right SMA showed relationships to out-of-scanner language measures that differed in direction between patients with larger vs. smaller lesions (Figure 5C). One explanation as to why stronger activations of these regions were associated with poorer performance for patients with smaller lesions is that activity in these regions interferes with the functions of task- relevant areas when canonical regions are intact, but supports residual language function when canonical regions are damaged. Alternately, stronger activations of these regions for patients with smaller lesions may be reflective of focal damage to critical left hemispheric areas that impede the restoration of canonical network function during recovery, and result in chronic language deficits comparable to those observed in patients with more widespread left hemispheric damage. Based on evidence that both the right IFG and right SMA support language processing during early recovery, but show reduced involvement in later stages when canonical networks are recovered [Saur et al., 2006], we consider this to be most likely. If this is the case, the identification of such critical regions by future studies could provide important insights into the factors contributing to poor long-term language recovery after stroke. Along these lines, we note that previous studies suggest that activation in the right IFG pars opercularis is increased in patients with lesions affecting the left IFG [Blank et al., 2003; Turkeltaub et al., 2011]. In addition, maintained right frontal activation has been previously reported in patients with left posterior temporal lesions that recovered less successfully than patients with lesions affecting the left frontal cortices or the left basal ganglia [Heiss et al., 1999]. Given that (1) our results indicate that larger lesions correlated with larger activation magnitudes in right frontal cortices, and (2) activation magnitudes in right IFG/SMA were associated with better expressive language functions in patients with larger lesions, we speculate that the observed benefit of activating the right IFG/right SMA in patients with larger lesions may reflect beneficial compensation that supports expressive language functions (e.g. top-down selection and sequencing). Based on the findings of Heiss and colleagues [1999], we further speculate that patients with lesions affecting left posterior temporal regions may also recruit these regions, but that these patients may show reduced benefit from their involvement, presumably resulting from an inability to restore function in other distributed portions of canonical networks. We stress that this must be regarded as speculative, as the current study did not directly investigate relationships between lesion location and right hemispheric activation. Rather, the methods used in this study were primarily intended to address the question of how lesion extent relates to right hemispheric activation. Because the methodology of this study was not ideal for addressing questions about how lesion location relates to regional activation, future studies using methodology better suited to addressing this question are necessary before strong conclusions can be drawn about how the site of damage relates to right hemispheric activations. A common network supporting comprehension, naming, and fluency Activity in a subset of regions (shown in Figure 3B) correlated positively with performance on all of the language measures, suggesting that this set of regions may support processes critical for language processing after stroke. This set of regions included the left dorsal dSFG, the left AG/superior lateral occipital gyrus, bilateral precuneus and PCC, bilateral PHG and fusiform gyri, and the right temporal pole. These regions are commonly associated with the default mode network [Dosenbach et al., 2007; Fox et al., 2005; Raichle et al., 2001; Raichle and Snyder, 2007; Vincent et al., 2008]. Based on observations that this network is more active during both semantic processing tasks and task-free resting states relative to tasks that do not involve semantic processing (and generally more active during semantic processing than during rest), this network has been proposed to form a core "conceptual network" that is involved in both explicit semantic processing and ongoing manipulations of memory/conceptual representations in the absence of explicit tasks [Binder et al., 1999; Binder et al., 2009; Leech and Sharp, 2014]. Notably, the PCC, precuneus, and AG are functionally connected to the PHG and co-activate with parahippocampal areas during memory retrieval [Sestieri et al., 2011]. The AG [Uddin et al., 2010] and PCC [Greicius et al., 2009] also possess direct structural connections to medial temporal lobe structures such as the hippocampus that play a critical role in memory. Speculatively, since each of the language measures utilized in this study involved a memory component (e.g. recalling facts about animals for the in- scanner semantic decision task, recalling words that begin with a given letter or fit a given category for the COWAT/SFT, and recalling names of visual objects for the BNT), it is possible that activity in these regions reflects their role in supporting memory access and manipulation. In addition, recent intrinsic functional connectivity evidence suggests that regions such as the left dSFG and left AG may function as hubs that support ongoing interactions among canonical language modules (i.e. perisylvian language areas), executive control modules (i.e. fronto-parietal network), and memory/simulation modules (i.e. default-mode network) [Xu et al., 2016]. Additionally, the right anterior temporal lobe has previously been reported to support auditory verbal comprehension in patients with damage to left posterior temporal areas, and may represent an independent module capable of auditory language processing when left posterior temporal structures are compromised [Crinion and Price, 2005]. When connectivity between left and right anterior temporal areas is preserved, left frontal access to inputs processed by the right anterior temporal lobe may provide a compensatory mechanism for achieving top-down modulations (e.g. selection or integration) of semantic content via inter-hemispheric pathways [Warren et al., 2009]. In addition, medial posterior default mode regions frequently interact with elements of other functional networks [de Pasquale et al., 2012] and possess diverse structural connections to language-relevant cortical (i.e. inferior parietal, superior temporal, anterior cingulate, and supplementary/pre-motor cortices) and subcortical (including the striatum and multiple thalamic nuclei) areas whose connections may be disrupted by LMCA stroke [Cavanna and Trimble, 2006; Greicius et al., 2009]. Thus, this set of regions could potentially act to relay information among surviving language areas when primary pathways are affected by stroke, providing a potential "back-up" interface among bilateral frontal, temporal, and parietal areas to support manipulations of memory representations and/or speech inputs when canonical cortico-cortical pathways (e.g. the arcuate fasciculus) are no longer viable. While plausible, and consistent with a proposed role of these regions as cross-network "connectors" [Xu et al., 2016], this should be regarded as speculation, as delineating the precise roles these regions play in supporting language processing after stroke is beyond the scope of this study. Lesion volume effects in functional neuroimaging of aphasia recovery Because we used statistical controls to account for lesion volume effects, our results are not likely to be driven by differences in lesion size or overall damage to the network among patients. Despite the potential for lesion volume differences to introduce bias, we note that our results were nearly identical (albeit with somewhat larger effects) when we did not use controls for lesion volume. This may reflect the relatively low correlation between lesion volume and canonical network activation in this sample. Nonetheless, while we contend that the use of statistical controls is important to reduce the potential for biases related to differences in lesion extent, it is possible that lesion volume is too coarse of a measure to accomplish this goal. An alternative approach may be to utilize more spatially sensitive measures, such as the percent of damage sustained by different anatomical areas [e.g. Xing et al., 2015]. Determining the optimal solution to account for lesion biases in functional neuroimaging studies of stroke patients may ultimately require dedicated studies using simulated data to understand the severity of This study has several limitations that must be acknowledged. First, as with many biases likely to be introduced by varying levels of brain damage and to allow for an objective comparison of different approaches for mitigating such biases. Limitations functional imaging studies, it is not possible to definitively conclude that activation in any given region is necessary to perform the task. Further, even if this were possible, it would not be possible to conclude from these data which aspect of the task a hypothetical "task-critical" region might support. Thus, while our results are interpreted as indicating that canonical network regions likely support residual language abilities after left hemispheric stroke, and that specific out-of-network regions in the right hemisphere may compensate for extensive left hemispheric damage, conclusions about the nature of the contributions of these regions to task performance cannot be considered definitive. Second, while the current study suggests that extensive left hemispheric damage is associated with the recruitment of select right frontal regions to accomplish semantic processing, it cannot address important questions such as how damage to specific regions in the left hemisphere influences semantic processing or affects out-of-network responses in the right hemisphere; future research is necessary to address these questions. Third, altered vascular dynamics in chronic stroke patients have the potential to lead to changes in neurovascular coupling and influence the BOLD response, and this may lead to reductions in measured activation magnitudes for stroke patients relative to controls [e.g. D'Esposito et al., 2003; Veldsman et al., 2015]. It is necessary to consider the potential for such effects when interpreting the results of this and other studies using functional MRI to study clinical populations such as stroke patients. Fourth, this study was not intended to address the question of how lesion location relates to right hemispheric activation. We stress that strong conclusions regarding this relationship cannot be drawn from the results of this study. Finally, the current study does not allow for conclusions regarding the role of functional interactions among regions associated with the CSN, between left and right hemispheric portions of the CSN, or between the CSN and other networks in supporting language recovery after stroke, as it is limited to assessments of In conclusion, this study aimed to test current theories of language recovery after co-activation during task performance. Future studies using functional and/or effective connectivity measures are necessary to enable such conclusions. Conclusions stroke and bring clarity to discrepancies in the literature. Our results confirm the prediction that fMRI activation in the canonical semantic network predicts performance on multiple measures of language function in chronic patients independently of lesion volume effects, and reveal a core set of default mode regions that likely play a key role in supporting basic aspects of residual language functions in the years after stroke. Our results also suggest that left and right hemispheric portions of the semantic network co- activate to accomplish language processing after stroke, contrary to the notion that activity in right hemispheric homologues of the left hemispheric network is up-regulated to compensate for damage to the left hemispheric network. In contrast, our results suggest that patients with the most damage to the left hemispheric network recruit specific right hemispheric areas that have been previously shown to play a primary role in supporting early recovery. We suspect that these regions may support top-down selection or motor aspects of residual language function during the chronic stage in these patients. The findings described here emphasize the importance of canonical language networks for supporting long-term language recovery after stroke, clarify the contributions of in- and out- of network right hemispheric areas to language recovery and their relationships to left hemisphere damage, and provide a more stable foundation for future studies of language recovery after stroke. Acknowledgements Amber Martin Christi Banks NIH R01 HD068488 NIH R01 NS048281 References Aickin M, Gensler H (1996): Adjusting for multiple testing when reporting research results: The Bonferroni vs Holm methods. Am J Public Health 86:726–728. Allendorfer J, Kissela B, Holland S, Szaflarski J (2012): Different patterns of language activation in post-stroke aphasia are detected by overt and covert versions of the verb generation fMRI task. Med Sci Monit Monit 18. Benjamini Y, Hochberg Y (1995): Controlling the False Discovery Rate: A Practical and Powerful Approach to Multiple Testing. J R Stat Soc Ser B 57:289–300. Baldassarre A, Ramsey L, Rengachary J, Zinn K, Siegel JS, Metcalf N V., Strube MJ, Snyder AZ, Corbetta M, Shulman GL (2016): Dissociated functional connectivity profiles for motor and attention deficits in acute right-hemisphere stroke. Brain 139:2024–2038. Belin P, Van Eeckhout P, Zilbovicius M, Remy P, François C, Guillaume S, Chain F, Rancurel G, Samson Y (1996): Recovery from nonfluent aphasia after melodic intonation therapy: a PET study. Neurology 47:1504–1511. Binder JR, Frost JA, Hammeke TA, Cox RW, Rao SM, Prieto T (1997): Human brain language areas identified by functional magnetic resonance imaging. J Neurosci 17:353–362. Binder JR (2012): Task-induced deactivation and the "resting" state. Neuroimage 62:1086–1091. Binder JR, Desai RH, Graves WW, Conant LL (2009): Where is the semantic system? A critical review and meta-analysis of 120 functional neuroimaging studies. Cereb Cortex 19:2767–2796. Binder JR, Swanson SJ, Hammeke TA, Sabsevitz DS (2008): A comparison of five fMRI protocols for mapping speech comprehension systems. Epilepsia 49:1980–1997. Binder J, Frost J, Hammeke T, Bellgowan PSF, Rao SM, Cox RW (1999): Conceptual processing during the conscious resting state: A functional MRI study. J Cogn Neurosci 11:80–93. Blank SC, Bird H, Turkheimer F, Wise RJS (2003): Speech production after stroke: The role of the right pars opercularis. Ann Neurol 54:310–320. Butler RA, Lambon Ralph MA, Woollams AM (2014): Capturing multidimensionality in stroke aphasia: mapping principal behavioural components to neural structures. Brain:3248–3266. Carter AR, Shulman GL, Corbetta M (2012): Why use a connectivity-based approach to study stroke and recovery of function? Neuroimage 62:2271–2280. Cavanna AE, Trimble MR (2006): The precuneus: a review of its functional anatomy and behavioural correlates. Brain 129:564–83. Charidimou A, Kasselimis D, Varkanitsa M, Selai C, Potagas C, Evdokimidis I (2014): Why is it difficult to predict language impairment and outcome in patients with aphasia after stroke? J Clin Neurol 10:75–83. Cheng B, Forkert ND, Zavaglia M, Hilgetag CC, Golsari A, Siemonsen S, Fiehler J, Pedraza S, Puig J, Cho T-H, Alawneh J, Baron J-C, Ostergaard L, Gerloff C, Thomalla G (2014): Influence of Stroke Infarct Location on Functional Outcome Measured by the Modified Rankin Scale. Stroke 45:1695–1702. Cole MW, Bassett DS, Power JD, Braver TS, Petersen SE (2014): Intrinsic and task- evoked network architectures of the human brain. Neuron 83:238–251. D'Esposito M, Deouell LY, Gazzaley A (2003): Alterations in the BOLD fMRI signal with ageing and disease: a challenge for neuroimaging. Nat Rev Neurosci 4:863– 872. Donnelly KM, Allendorfer JB, Szaflarski JP (2011): Right hemispheric participation in semantic decision improves performance. Brain Res 1419:105–116. Dosenbach NUF, Visscher KM, Palmer ED, Miezin FM, Wenger KK, Kang HC, Burgund ED, Grimes AL, Schlaggar BL, Petersen SE (2006): A Core System for the Implementation of Task Sets. Citeulike:JOUR. Neuron 50:799–812. Dosenbach NU, Fair DA, Miezin FM, Cohen AL, Wenger KK, Dosenbach RAT, Fox MD, Snyder AZ, Vincent JL, Raichle ME, Schlaggar BL, Petersen SE (2007): Distinct brain networks for adaptive and stable task control in humans. Proc Natl Acad Sci U S A 104:11073–8. Eaton KP, Szaflarski JP, Altaye M, Ball AL, Kissela BM, Banks C, Holland SK (2008): Reliability of fMRI for studies of language in post-stroke aphasia subjects. Neuroimage 41:311–22. Eklund A, Nichols TE, Knutsson H (2016): Cluster failure: Why fMRI inferences for spatial extent have inflated false-positive rates. Proc Natl Acad Sci 113:201602413. van Ettinger-Veenstra HM, Ragnehed M, Hällgren M, Karlsson T, Landtblom A-M, Lundberg P, Engström M (2010): Right-hemispheric brain activation correlates to language performance. Neuroimage 49:3481–8. van Ettinger-Veenstra H, Ragnehed M, McAllister A, Lundberg P, Engström M (2012): Right-hemispheric cortical contributions to language ability in healthy adults. Brain Lang 120:395–400. Fisher R (1921): On the probable error of a coefficient of correlation deduced from a small sample. Metron 1:3–32. Fox MD, Corbetta M, Snyder AZ, Vincent JL, Raichle ME (2006): Spontaneous neuronal activity distinguishes human dorsal and ventral attention systems. Proc Natl Acad Sci U S A 103:10046–51. Fox MD, Snyder AZ, Vincent JL, Corbetta M, Van Essen DC, Raichle ME (2005): The human brain is intrinsically organized into dynamic, anticorrelated functional networks. Proc Natl Acad Sci U S A 102:9673–8. Fridriksson J, Bonilha L, Baker JM, Moser D, Rorden C (2010): Activity in preserved left hemisphere regions predicts anomia severity in aphasia. Cereb Cortex 20:1013–9. Fridriksson J, Richardson JD, Fillmore P, Cai B (2012): Left hemisphere plasticity and aphasia recovery. Neuroimage 60:854–63. Friston KJ, Holmes AP, Worsley KJ, Poline J-P, Frith CD, Frackowiak RSJ (1995): Statistical parametric maps in functional imaging: A general linear approach. Hum Brain Mapp 2:189–210. Geranmayeh F, Leech R, Wise RJS (2016): Network dysfunction predicts speech production after left hemisphere stroke. Neurology 86:1296–1305. Griffis JC, Allendorfer JB, Szaflarski JP (2016a): Voxel-based Gaussian naïve Bayes classification of ischemic stroke lesions in individual T1-weighted MRI scans. J Neurosci Methods 257:97–108. Griffis JC, Nenert R, Allendorfer JB, Szaflarski JP (2016b): Interhemispheric Plasticity following Intermittent Theta Burst Stimulation in Chronic Poststroke Aphasia. Neural Plast 2016:20–23. Greicius MD, Supekar K, Menon V, Dougherty RF (2009): Resting-state functional connectivity reflects structural connectivity in the default mode network. Cereb Cortex 19:72–8. Hamilton RH, Chrysikou EG, Coslett B (2011): Mechanisms of aphasia recovery after stroke and the role of noninvasive brain stimulation. Brain Lang 118:40–50. Hartwigsen G, Saur D, Price CJ, Ulmer S, Baumgaertner A, Siebner HR (2013): Perturbation of the left inferior frontal gyrus triggers adaptive plasticity in the right homologous area during speech production. Proc Natl Acad Sci U S A 110:16402–7. He BJ, Snyder AZ, Vincent JL, Epstein A, Shulman GL, Corbetta M (2007): Breakdown of functional connectivity in frontoparietal networks underlies behavioral deficits in spatial neglect. Neuron 53:905–18. Heiss WD, Kessler J, Thiel A, Ghaemi M, Karbe H (1999): Differential capacity of left and right hemispheric areas for compensation of poststroke aphasia. Ann Neurol 45:430–438. Heiss W, Thiel A (2006): A proposed regional hierarchy in recovery of post-stroke aphasia. Brain Lang 98:118–23. Jaccard J, Wan CK, Turrisi R (1990): The Detection and Interpretation of Interaction Effects Between Continuous Variables in Multiple Regression. Multivariate Behav Res 25:467–478. Kaplan E, Goodglass H, Weintraub S, Segal O, van Loon-Vervoorn A (2001): Boston naming test. Pro-ed. Karbe H, Thiel A, Weber-luxenburger G, Kessler J, Heiss W (1998): Brain Plasticity in Poststroke Aphasia  : What Is the Contribution of the Right Hemisphere? Brain Lang 230:215–230. Karnath HO, Berger MF, Kuker W, Rorden C (2004): The anatomy of spatial neglect based on voxelwise statistical analysis: A study of 140 patients. Cereb Cortex 14:1164–1172. Kertesz A, Harlock W, Coates R (1979): Computer tomographic localization, lesion size, and prognosis in aphasia and nonverbal impairment. Brain Lang 8:34–50. Kim KK, Karunanayaka P, Privitera MD, Holland SK, Szaflarski JP (2011): Semantic association investigated with functional MRI and independent component analysis. Epilepsy Behav 20:613–622. Kozora E, Cullum CM (1995): Generative naming in normal aging: Total output and qualitative changes using phonemic and semantic constraints. Clin Neuropsychol 9:313–320. Kümmerer D, Hartwigsen G, Kellmeyer P, Glauche V, Mader I, Klöppel S, Suchan J, Karnath HO, Weiller C, Saur D (2013): Damage to ventral and dorsal language pathways in acute aphasia. Brain 136:619–629. Lazar RM, Speizer AE, Festa JR, Krakauer JW, Marshall RS (2008): Variability in language recovery after first-time stroke. J Neurol Neurosurgery, Psychiatry 79:530–534. Leech R, Sharp DJ (2014): The role of the posterior cingulate cortex in cognition and disease. Brain 137:12–32. Lezak MD, Howieson DB, Loring DW, Hannay JH, Fischer JS (1995): Neuropsychological assessment (3) Oxford University Press. New York. Maas MB, Lev MH, Ay H, Singhal AB, Greer DM, Smith WS, Harris GJ, Halpern EF, Koroshetz WJ, Furie KL (2012): The Prognosis for Aphasia in Stroke. J Stroke Cerebrovasc Dis 21:350–357. Mattioli F, Ambrosi C, Mascaro L, Scarpazza C, Pasquali P, Frugoni M, Magoni M, Biagi L, Gasparotti R (2014): Early aphasia rehabilitation is associated with functional reactivation of the left inferior frontal gyrus: a pilot study. Stroke 45:545– 52. Mazaika PK, Whitfield S, Cooper JC (2005): Detection and repair of transient artifacts in fMRI data. Neuroimage 26:S36. Meinzer M, Beeson PM, Cappa S, Crinion J, Kiran S, Saur D, Parrish T, Crosson B, Thompson CK (2013): Neuroimaging in aphasia treatment research: consensus and practical guidelines for data analysis. Neuroimage 73:215–24. Meltzer JA, Wagage S, Ryder J, Solomon B, Braun AR (2013): Adaptive significance of right hemisphere activation in aphasic language comprehension. Neuropsychologia 51:1248–1259. Muhle-Karbe PS, Derrfuss J, Lynn MT, Neubert FX, Fox PT, Brass M, Eickhoff SB (2015): Co-Activation-Based Parcellation of the Lateral Prefrontal Cortex Delineates the Inferior Frontal Junction Area. Cereb Cortex:1–17. van Oers CMM, Vink M, van Zandvoort MJE, van der Worp HB, de Haan EHF, Kappelle LJ, Ramsey NF, Dijkhuizen RM (2010): Contribution of the left and right inferior frontal gyrus in recovery from aphasia. A functional MRI study in stroke patients with preserved hemodynamic responsiveness. Neuroimage 49:885–93. Oldfield RC (1971): the Assessment and Analysis of Handedness: the Edinburgh Inventory. Neuropsychologia 9:97–113. Ovadia-Caro S, Villringer K, Fiebach J, Jungehulsing GJ, van der Meer E, Margulies DS, Villringer A (2013): Longitudinal effects of lesions on functional networks after stroke. J Cereb Blood Flow Metab 33:1279–85. Pani E, Zheng X, Wang J, Norton A, Schlaug G (2016): Right hemisphere structures predict poststroke speech fluency. Neurology 86:1574–1581. de Pasquale F, Della Penna S, Snyder AZ, Marzetti L, Pizzella V, Romani GL, Corbetta M (2012): A cortical core for dynamic integration of functional networks in the resting human brain. Neuron 74:753–64. Pedersen P (1995): Aphasia in acute stroke: incidence, determinants, and recovery. Ann Neurol 38:659–666. Piervincenzi C, Petrilli A, Marini A, Caulo M, Committeri G, Sestieri C (2016): Multimodal assessment of hemispheric lateralization for language and its relevance for behavior. Neuroimage. Poldrack RA (2012): The future of fMRI in cognitive neuroscience. Neuroimage 62:1216–20. Raboyeau G, De Boissezon X, Marie N, Balduyck S, Puel M, Bézy C, Démonet JF, Cardebat D (2008): Right hemisphere activation in recovery from aphasia: Lesion effect or function recruitment? Neurology 70:290–298. Raichle ME, MacLeod AM, Snyder AZ, Powers WJ, Gusnard D a, Shulman GL (2001): A default mode of brain function. Proc Natl Acad Sci U S A 98:676–82. Raichle ME, Snyder AZ (2007): A default mode of brain function: a brief history of an evolving idea. Neuroimage 37:1083–90; discussion 1097–9. Ripollés P, Marco-Pallarés J, de Diego-Balaguer R, Miró J, Falip M, Juncadella M, Rubio F, Rodriguez-Fornells A (2012): Analysis of automated methods for spatial normalization of lesioned brains. Neuroimage 60:1296–306. Rorden C, Karnath H-O (2004): Using human brain lesions to infer function: a relic from a past era in the fMRI age? Nat Rev Neurosci 5:813–819. Rosen HJ, Petersen SE, Linenweber MR, Snyder AZ, White DA, Chapman L, Dromerick AW, Fiez JA, Corbetta MD (2000): Neural correlates of recovery from aphasia after damage to left inferior frontal cortex. Neurology 55:1883–1894. Saur D, Hartwigsen G (2012): Neurobiology of language recovery after stroke: lessons from neuroimaging studies. Arch Phys Med Rehabil 93:S15-25. Saur D, Lange R, Baumgaertner A, Schraknepper V, Willmes K, Rijntjes M, Weiller C (2006): Dynamics of language reorganization after stroke. Brain 129:1371–84. Schwartz MF, Kimberg DY, Walker GM, Faseyitan O, Brecher A, Dell GS, Coslett HB (2009): Anterior temporal involvement in semantic word retrieval: voxel-based lesion-symptom mapping evidence from aphasia. Brain 132:3411–27. Seghier ML, Ramlackhansingh A, Crinion J, Leff AP, Price CJ (2008): Lesion identification using unified segmentation-normalisation models and fuzzy clustering. Neuroimage 41:1253–1266. Sestieri C, Corbetta M, Romani GL, Shulman GL (2011): Episodic memory retrieval, parietal cortex, and the default mode network: functional and topographic analyses. J Neurosci 31:4407–20. Siegel JS, Ramsey LE, Snyder AZ, Metcalf N V, Chacko RV, Weinberger K, Baldassarre A, Hacker C, Shulman GL (2016): Common and specific disruptions of network connectivity predict impairment in multiple behavioral domains after stroke. PNAS I:1–10. Sims JA, Kapse K, Glynn P, Sandberg C, Tripodis Y, Kiran S (2016): The relationships between the amount of spared tissue, percent signal change, and accuracy in semantic processing in aphasia. Neuropsychologia 84:113–126. Slotnick SD, Moo LR, Segal JB, Hart J (2003): Distinct prefrontal cortex activity associated with item memory and source memory for visual shapes. Cogn Brain Res 17:75–82. Szaflarski JP, Binder JR, Possing ET, McKiernan KA, Ward BD, Hammeke TA (2002): Language lateralization in left-handed and ambidextrous people: fMRI data. Neurology 59:238–244. Szaflarski JP, Allendorfer JB, Banks C, Vannest J, Holland SK (2013): Recovered vs. not-recovered from post-stroke aphasia: the contributions from the dominant and non-dominant hemispheres. Restor Neurol Neurosci 31:347–60. Szaflarski JP, Holland SK, Jacola LM, Lindsell C, Privitera MD, Szaflarski M (2008): Comprehensive presurgical functional MRI language evaluation in adult patients with epilepsy. Epilepsy Behav 12:74–83. Szaflarski JP, Vannest J, Wu SW, DiFrancesco MW, Banks C, Gilbert DL (2011): Excitatory repetitive transcranial magnetic stimulation induces improvements in chronic post-stroke aphasia. Med Sci Monit 17:CR132-9. Thiel A, Schumacher B, Wienhard K, Gairing S, Kracht LW, Wagner R, Haupt WF, Heiss W-D (2006): Direct demonstration of transcallosal disinhibition in language networks. J Cereb Blood Flow Metab 26:1122–7. Turkeltaub PE, Coslett HB, Thomas AL, Faseyitan O, Benson J, Norise C, Hamilton RH (2012): The right hemisphere is not unitary in its role in aphasia recovery. Cortex 48:1179–1186. Turkeltaub PE, Messing S, Norise C, Hamilton RH (2011): Are networks for residual language function and recovery consistent across aphasic patients? Neurology 76:1726–34. Uddin LQ, Supekar K, Amin H, Rykhlevskaia E, Nguyen DA, Greicius MD, Menon V (2010): Dissociable connectivity within human angular gyrus and intraparietal sulcus: evidence from functional and structural connectivity. Cereb Cortex 20:2636– 46. Veldsman M, Cumming T, Brodtmann A (2015): Beyond BOLD: Optimizing functional imaging in stroke populations. Hum Brain Mapp 36:1620–1636. Vincent JL, Kahn I, Snyder AZ, Raichle ME, Buckner RL (2008): Evidence for a frontoparietal control system revealed by intrinsic functional connectivity. J Neurophysiol 100:3328–42. Wager TD, Lindquist M, Kaplan L (2007): Meta-analysis of functional neuroimaging data: Current and future directions. Soc Cogn Affect Neurosci 2:150–158. Warren JE, Crinion JT, Lambon Ralph MA, Wise RJS (2009): Anterior temporal lobe connectivity correlates with functional outcome after aphasic stroke. Brain 132:3428–42. Wilke M, de Haan B, Juenger H, Karnath H-O (2011): Manual, semi-automated, and automated delineation of chronic brain lesions: a comparison of methods. Neuroimage 56:2038–46. Winhuisen L, Thiel A, Schumacher B, Kessler J, Rudolf J, Haupt WF, Heiss WD (2005): Role of the contralateral inferior frontal gyrus in recovery of language function in poststroke aphasia: a combined repetitive transcranial magnetic stimulation and positron emission tomography study. Stroke 36:1759–63. Winhuisen L, Thiel A, Schumacher B, Kessler J, Rudolf J, Haupt WF, Heiss WD (2007): The right inferior frontal gyrus and poststroke aphasia: a follow-up investigation. Stroke 38:1286–92. Xing S, Lacey EH, Skipper-Kallal LM, Jiang X, Harris-Love ML, Zeng J, Turkeltaub PE (2015): Right hemisphere grey matter structure and language outcomes in chronic left hemisphere stroke. Brain. Xu Y, Lin Q, Han Z, He Y, Bi Y (2016): Intrinsic functional network architecture of human semantic processing: Modules and hubs. Neuroimage 132:542–555. Yarkoni T (2009): Big Correlations in Little Studies: Inflated fMRI Correlations Reflect Low Statistical Power-Commentary on Vul et al. (2009). Perspect Psychol Sci 4:294–298. Yarnell P, Monroe P, Sobel L (1976): Aphasia outcome in stroke: a clinical neuroradiological correlation. Stroke 7:516–522. Zhang Y, Kimberg D, Coslett H, Schwartz M, Wang Z (2014): Multivariate lesion- symptom mapping using support vector regression. Hum Brain Mapp 35:997. Tables Table 1. Participant demographics Group N Age Patients 43 53 (15) 25 M 0.85 (0.43) Controls 43 54( 14) 23 M 0.80 (0.41) *Mean (SD) are shown for age/handedness; M-Male Table 2. Cluster and peak statistics for control and patient SD activations Sex Handedness Lesion volume (ml) 105.24 (76.29) N/A x -4 2 Peak Location y z Extent t-value 46 36 44277 14.3308 -34 32 44277 10.8963 -8 44277 10.6975 -36 30 3981 11.8496 -40 -64 36 -60 32 1051 50 125 -12 66 2 -76 -28 3519 10 3519 -54 -52 8 3519 -66 -48 42 7349 40 56 -2 -48 28 7349 8 7349 16 54 30 253 -52 -8 -18 -44 -64 28 1085 -68 32 254 54 217 0 -32 30 -60 -32 -12 585 111 40 -24 -28 265 4 -46 14 Group Controls L Superior Medial Gyrus R Posterior Cingulate L IFG (p. Orbitalis) L Angular Gyrus R Angular Gyrus R Mid Orbital Gyrus Patients R Cerebelum (VIII) R Cerebelum (IX) R Cerebelum (VII) L Superior Medial Gyrus L IFG (p. Triangularis) R Superior Frontal Gyrus L Middle Temporal Gyrus L Middle Temporal Gyrus R Angular Gyrus Posterior Cingulate L Inferior Temporal Gyrus R Fusiform Gyrus R Posterior Cingulate *Note: Cluster peaks are provided for the top 3 peaks/cluster that are separated by a minimum distance of 30mm. Table 3. Cluster and peak statistics for correlations between patient SD activation and language task performance while controlling for lesion volume. y 8.1312 3.3213 5.1981 4.1282 3.8702 4.9037 4.7172 4.0744 4.409 4.4082 4.0334 3.9907 3.9528 3.7833 2.9714 Peak Location t-value x z Sign Positive L Angular Gyrus L Middle Temporal Gyrus Cerebellar Vermis (VII) Posterior Cingulate Gyrus L Calcarine Gyrus Dorsal Anterior Pons L Middle Orbital Gyrus L Middle Temporal Gyrus L Inferior Temporal Gyrus R Angular Gyrus L Parahippocampal Gyrus R Parahippocampal Gyrus L Superior Frontal Gyrus R IFG (p. Orbitalis) R Precuneus Extent Semantic Decisions 1913 1913 5291 5291 5291 544 961 785 671 547 457 275 175 160 211 -42 -68 48 4.9167 -50 -46 8 2.5892 -2 -78 -18 4.8921 0 -32 32 4.4726 -94 12 4 4.3329 0 -12 -28 4.6492 -42 54 0 4.5543 -2 4.4425 -68 -32 -32 -6 -38 4.3857 -58 40 4.2227 54 -22 -28 -18 4.0258 3.9305 20 -32 -20 3.745 -20 56 32 3.7179 50 38 -18 -62 66 4 3.6912 7588 Positive R Superior Medial Gyrus 7588 L Middle Frontal Gyrus 7588 L Posterior-Medial Frontal L Brainstem (Inf. Colliculus?) 6302 L Fusiform Gyrus 6302 R Cerebelum (IX) 6302 1227 L Angular Gyrus L Posterior Cingulate 4805 R Cerebelum (VI) 4805 4805 L Calcarine Gyrus R Fusiform Gyrus 232 R Medial Temporal Pole 1532 1532 R Middle Temporal Gyrus 1532 R Parahippocampal Gyrus L IFG (p. Orbitalis) 587 Negative R Middle Temporal Gyrus 155 Naming 2 50 38 6.0189 -34 18 52 5.8247 -4 20 66 5.1085 -2 -30 -20 5.2447 -34 -18 -30 4.9057 8 -56 -38 4.8778 -48 -62 32 5.2081 -2 -46 18 4.6649 -80 -14 4.0471 28 6 -92 3.6055 0 4.616 34 -18 -34 4.5911 46 6 -36 -16 -20 3.7147 68 -30 -2 3.685 16 4.0197 -12 -40 24 -3.3541 48 -48 2 Negative L Postcentral Gyrus R Rectal Gyrus R Inferior Temporal Gyrus R Temporal Pole 135 153 421 258 3.6875 2 24 -30 3.6576 56 -12 -30 3.6249 36 14 -24 -3.4808 -60 -10 30 Fluency 5204 5204 5204 1889 268 474 354 265 265 762 2516 2516 2516 202 139 54 46 2 5.5202 -34 16 52 4.494 14 74 6 3.4022 -46 -60 30 4.2896 48 2 -34 4.207 -40 -34 -14 3.9476 44 24 -14 3.921 -42 3.9092 16 -8 2.6356 16 14 -18 -46 -38 3.8878 14 3.8798 -66 64 2 3.5849 0 -36 32 -28 -56 66 2.4993 -42 26 3.8209 -16 3.6031 -18 -38 6 Positive R Superior Medial Gyrus L Middle Frontal Gyrus R Posterior-Medial Frontal L Angular Gyrus R Inferior Temporal Gyrus L Fusiform Gyrus R IFG (p. Orbitalis) R Fusiform Gyrus R Superior Orbital Gyrus R Cerebelum (IX) R Precuneus Posterior Cingulate L Superior Parietal Lobule L IFG (p. Orbitalis) L Fusiform Gyrus Note: Cluster peaks are provided for the top 3 peaks/cluster that are separated by a minimum distance of 30mm Table 4. Cluster and peak statistics for right hemispheric correlations between SD activation and lesion volume. Sign x Extent 277 200 143 221 399 223 1137 1137 1137 160 Peak Location Positive R IFG (p. Opercularis) R Posterior-Medial Frontal R Middle Frontal Gyrus R Precentral Gyrus Negative R Inferior Temporal Gyrus R Mid Orbital Gyrus R Cerebelum (Crus I) R Cerebelum (IX) R Cerebelum (Crus I) R Anterior Thalamus/Putamen Note: Cluster peaks are provided for the top 3 peaks/cluster that are separated by a minimum distance of 30mm Figures z t-value y 8 3.9446 56 12 8 56 3.0469 6 3.0383 34 40 20 40 -2 2.9707 54 -4.1944 56 -14 -34 -3.9815 2 46 4 -3.8737 56 -62 -32 -3.3749 14 -56 -48 -3.1767 34 -82 -24 -3.7215 10 -6 -2 Figure 1. Sample sizes in the functional neuroimaging literature of post-stroke aphasia. Histogram illustrating frequencies of different sample sizes across 84 functional neuroimaging studies of aphasia published since 1995. Bins correspond to single integer values. The average sample size was 10.33, with a standard deviation of 8.44. The single largest sample was 53 patients. 67 (81%) of studies had samples sizes less than or equal to 15 patients, and 73 (87%) of studies had sample sizes less than or equal to 20 patients. Individual study sample sizes and neuroimaging modalities are provided in Supplementary Table 1, and the full reference list is also provided in Supplementary Material 1. Figure 2. Data characterization and partial behavioral correlation results. A. Lesion frequencies across all 43 patients. Minimum colorbar values indicate voxels lesioned in only 1 patient, and maximum colorbar values indicate voxels lesioned in 32 patients (maximum lesion overlap). B. Means and standard errors for stroke patients and healthy controls are shown for each language measure. C. Areas showing significantly more activity during the SD condition relative to the control condition in the healthy controls (left) and areas showing significantly more activity during the SD condition relative to the TD condition in the stroke patients (right). Both maps are intensity thresholded at p<0.01, uncorrected and cluster-corrected at p<0.05 (99 voxels). Colorbar values indicate t-statistics. Note that the canonical semantic network (CSN) region of interest (ROI) was defined based on the results shown in (C, left). D. Scatterplots illustrate the partial correlations between the residuals for activation in the CSN ROI and the residuals for performance on each language measure (y-axes) after removing the effects of lesion volume for the patient group (top). Scatterplots illustrate correlations between activation in the CSN ROI and performance on each language measure for the control group (bottom). The line of best fit (red) and 95% confidence intervals (red dashes) are shown on each plot. Each relationship shown was significant at p<0.05, family-wise error corrected. Note: All brain renderings are in neurological convention. Figure 3. Whole-brain performance regression results. A. Positive correlations (hot colors) and negative correlations (cold colors) identified for the patient group between task-driven activation and performance on the SD task (top) combined fluency measure (middle) and Boston Naming Test (bottom). B. The subset of regions showing positive correlations between SD activation and performance on all language measures in the patient group are shown in red. C. The canonical network identified in controls is shown in green, regions within the CSN identified in healthy controls where SD activation was positively related to performance on any language measure in patients are shown in red, and regions outside of the CSN where SD activation was positively related to performance on any language measure in patients are shown in blue. Note that the overlays shown in (C) are qualitative illustrations of how the quantitatively identified behavioral relationships in patients relate to the canonical network identified in controls, and are intended to illustrate how activity supporting residual language task performance relates to the CSN. Each map is intensity thresholded at p<0.01, uncorrected and cluster- corrected at p<0.01 (126 voxels). Colorbar values indicate t-statistics. All brain renderings are in neurological convention. Figure 4. Relationships between left and right hemispheric activation. A. Brain renderings show the ROIs corresponding to the left hemispheric CSN (LH CSN), right hemispheric CSN (RH CSN), mirrored left hemispheric CSN (Mirror LH CSN), and out- of-network left hemispheric homologues (OON-LH-Hom). B. Scatterplots illustrating the relationship between residual (with the effects of lesion volume partialled out) SD-driven activity in the LH CSN ROI (y-axes) and each right hemispheric ROI (x-axes) are shown for the stroke group (top). Scatterplots illustrating the relationship between SD-driven activity in the LH CSN ROI (y-axes) and each right hemispheric ROI (x-axes) are also shown for the control group (bottom). C. Line plots illustrate relative activation magnitudes in the LH CSN ROI and each right hemispheric ROI for the stroke (left, red) and control (middle, blue) groups. Means and standard errors of activation magnitudes at each ROI are also shown for both groups (right). D. Scatterplots illustrate the relationships between left hemispheric lesion volume and activity in each ROI. Figure 5. Lesion effects on regional activation in the right hemisphere. A. Positive correlations (hot colors) and negative correlations (cold colors) between task-driven activation and lesion volume in right hemispheric areas are shown (top). B. Areas of overlap (red) and no overlap (blue) between regions where increased activation is associated with larger left hemispheric lesion volumes and the right hemispheric CSN ROI are shown on the top row. Areas of overlap (red) and no overlap (blue) between regions where decreased activation is associated with larger left hemispheric lesion volumes and the right hemispheric CSN ROI are shown on the middle row. Areas of overlap (red) and no overlap (blue) between regions where increased activation is associated with larger left hemispheric lesion volumes and the out-of-network left hemispheric homologues (OON-LH-Hom -- green) ROI are shown on the bottom row. Note that the overlays shown in (B) are qualitative illustrations of the quantitatively identified lesion volume effects on right hemispheric activations, and illustrate how right hemispheric activations associated with lesion volume effects relate to the RH CSN ROI (top, middle) and to the OON-LH-Hom ROI (bottom). C. Interaction plots are shown for the moderating effects of lesion volume on the relationship between right inferior frontal gyrus (IFG) activation and fluency scores (left), right supplementary motor area (SMA) activation and fluency scores (middle), and right SMA activation and naming scores (right). Each of these relationships survived a False Discovery Rate threshold of 0.1. High and low values shown in the interaction plots correspond to +1 and -1 standard deviations from the mean of each variable. All maps are intensity thresholded at p<0.01, uncorrected and cluster-corrected at p<0.01 (126 voxels). Colorbar values indicate t- statistics. Note: all brain renderings are in neurological convention. Study Study PET Abutalebi et al., 2009 PET Breier et al., 2009 PET Fridriksson et al., 2009 PET Martin et al., 2009 EEG Specht et al., 2009 NIRS Warren et al., 2009 PET PET fMRI PET PET Supplementary Material Supplementary Analysis 1. Sample sizes from literature search Supplementary Table 1. Aphasia patient sample sizes in 84 functional neuroimaging studies of aphasia between 1995 and 2016. N Modality N Modality Weiller et al., 1995 6 1 fMRI Belin et al., 1996 7 23 MEG Cappa et al., 1997 8 fMRI 11 Heiss et al., 1997 6 fMRI 2 Thomas et al., 1997 11 12 PET Sakatani et al., 1998 10 PET 24 Karbe et al., 1998 Fridriksson et al., 2010a 7 fMRI 15 Heiss et al., 1999 Fridriksson et al., 2010b 23 26 fMRI Miura et al., 1999 Postman-Caucheteux et al., 1 fMRI 3 2010 Musso et al., 1999 Rochon et al., 2010 4 4 fMRI Warburton et al., 1999 6 Saur et al., 2010 fMRI 21 Rosen et al., 2000 6 PET+fMRI Sharp et al., 2010 PET 9 Blasi et al., 2002 8 fMRI 5 Leger et al., 2002 1 6 fMRI Blank et al., 2003 14 fMRI 13 Cardebat et al., 2003 8 fMRI 14 Perani et al., 2003 5 4 fMRI Fernandez et al., 2004 1 fMRI 8 Kurland et al., 2004 2 fMRI 14 Naeser et al., 2004 4 10 PET Peck et al., 2004 3 fMRI 16 Zahn et al., 2004 7 30 fMRI Crinion et al., 2005 17 25 MEG Crosson et al., 2005 2 fMRI 27 De Boissezon et al., 7 fMRI 8 2005 Martin et al., 2005 5 24 PET Pulvermuller et al., 9 fMRI 14 2005 Winhuisen et al., 2005 11 16 fMRI Meinzer et al., 2006 1 fMRI 4 Connor et al., 2006 fMRI 12 8 Crinion et al., 2006 22 6 fMRI Fridriksson et al., 2006 3 fMRI 12 Saur et al., 2006 14 1 fMRI fMRI fMRI PET PET fMRI fMRI fMRI fMRI Weiduschat et al., 2011 Allendorfer et al., 2012 fMRI Fridriksson et al., 2012 fMRI fMRI Meltzer et al., 2013 Szaflarski et al., 2013 fMRI Thompson et al., 2013 PET fMRI Thiel et al., 2013 Abel et al., 2014 EEG PET Brownsett et al., 2014 Jarso et al., 2014 fMRI Mattioli et al., 2014 fMRI PET Mohr et al., 2014 Robson et al., 2014 fMRI fMRI Seghier et al., 2014 Thompson et al., 2010a Thompson et al., 2010b Van Oers et al., 2010 Papoutsi et al., 2011 Szaflarski et al., 2011a Szaflarski et al., 2011b Tyler et al., 2011 fMRI fMRI fMRI PET fMRI fMRI fMRI PET fMRI Bonakdarpour et al., 5 2007 Fridriksson et al., 2007 3 Meinzer et al., 2007 1 Winhuisen et al., 2007 9 Vitali et al., 2007 2 Eaton et al., 2008 4 Meinzer et al., 2008 11 Raboyeau et al., 2008 10 Richter et al., 2008 16 *References for supplementary Table 1 are at the end of the supplemental material. Supplementary Analysis 2. Additional characterization of patient data Van Hees et al., 2014 Zhu et al., 2014 Abel et al., 2015 Bonakdarpour et al., 2015 Kiran et al., 2015 Spironelli et al., 2015 Griffis et al., 2016 Sims et al., 2016 Geranmayeh et al., 2016 8 13 14 5 8 17 8 14 53 fMRI fMRI fMRI fMRI fMRI EEG fMRI fMRI fMRI Supplementary Figure 1. Behavioral measures. A. Plots of ordered scores for semantic decisions (left), averaged fluency (middle), and picture naming (right). Scores are shown separately for stroke patients (top, red) and controls (bottom, blue). B. Plots of relationships between patient scores on (1) naming and semantic decisions (left), (2) fluency and semantic decisions (middle), and (3) fluency and naming (right). For each plot, patient scores are shown as z-scores based on the control data. Supplementary Table 2. Patient characteristics Patient Age Sex EHI TSS BNT COWAT SFT Average Fluency %SD Correct 27 21 4 5 20 11 2 8 15 0 8 6 36 0 24 3 4 19 20 31 9 18 13 0 0 2 8 11 9 2 20 7 10 7 1 1 11 2 5 47 39 9 6 42 27 14 20 30 12 10 20 62 2 60 3 10 35 45 45 21 36 12 0 0 1 15 22 14 3 23 18 19 7 6 16 12 7 19 37 30 6.5 5.5 31 19 8 14 22.5 6 9 13 49 1 42 3 7 27 32.5 38 15 27 12.5 0 0 1.5 11.5 16.5 11.5 2.5 21.5 12.5 14.5 7 3.5 8.5 11.5 4.5 12 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 F 63 F 78 41 F 54 M 46 M 52 M 56 M 53 M 55 M 48 M 63 M 56 F 23 M 50 M F 48 70 F 68 M 59 M F 23 24 F 78 F 65 M F 58 72 F 50 M 57 M 51 M 43 M 24 M F 67 62 F 44 F 62 M 31 M 61 M 64 M 38 F 53 F 54 M 0.55 1.00 0.50 1.00 0.90 0.58 1.00 1.00 1.00 1.00 1.00 1.00 1.00 1.00 1.00 1.00 0.91 0.82 1.00 1.00 1.00 1.00 1.00 1.00 1.00 1.00 1.00 1.00 0.83 1.00 -1.00 0.91 1.00 1.00 1.00 -1.00 0.91 1.00 0.92 1.0 4.1 5.8 1.6 1.0 1.0 3.4 5.0 1.2 6.1 1.0 1.0 1.0 1.0 1.0 2.0 3.3 1.0 1.0 1.0 3.4 14.0 13.0 1.5 2.9 2.1 1.1 1.3 2.3 2.2 4.4 2.1 2.6 4.8 9.6 2.7 1.8 9.2 3.3 59 57 9 7 53 60 32 50 58 22 60 33 60 2 60 11 9 53 59 59 58 55 40 0 0 2 37 50 21 6 33 41 54 21 25 51 46 34 33 68 50 38 66 46 48 50 76 72 20 40 NA 76 28 72 0 50 72 76 58 36 68 36 NA 42 48 66 44 2 50 62 30 63 NA NA 30 72 74 55 1.00 0.71 1.00 1.00 1.3 1.3 3.4 12.4 31 3 50 34 1 0 4 2 46 M F 90 29 F 67 M 40 41 42 43 *EHI – Edinburgh Handedness Inventory, TSS – time since stroke, BNT – Boston Naming Test, COWAT – Controlled Oral Word Association Test, SFT – Semantic Fluency Test, %SD Correct -- % Semantic Decision Correct. Note: The average fluency score is the average of the COWAT and SFT. 5 0 11.5 9 9 0 19 16 34 0 0 26 Supplementary Figure 2. Patient lesions. Patient lesions are shown on example slices. Patients are organized in columns, as indicated by the text above each column. Supplementary Analysis 3. Evaluation of result dependence on use of lesion volume and scanner covariates additional analyses were performed to assess the stability of our results with and without scanner included as a covariate (since patients were collected on 2 different scanners) and with and without lesion volume as a covariate (where applicable). These analyses were performed to asses the effects of including scanner and lesion volume covariates on group results. Patients 1-20 were collected on the Phillips scanner, while patients 21-43 were collected on the Siemens scanner. We note that median lesion volumes significantly differed between patients collected on each scanner (Mann-Whitney U-test p=0.027), and mean lesion volumes showed a trend-level difference (t41 = 1.83, p=0.075) between scanners. Control participants were all collected on the Phillips scanner. Scanner and lesion volume covariates: Mean canonical semantic network (CSN) activation still predicted fluency (partial r =0.56, p<0.001), naming (partial r=0.49, p=0.001), and SD performance (partial r=0.46, p=0.006) when both scanner and lesion volume were partialled out. Activity in left and right hemispheric portions of the CSN remained significantly correlated when both lesion volume and scanner were partialled out (partial r=0.79, p<0.001). The same was true when the mirror ROI was used (partial r=0.80, p<0.001), and when the out-of-network homologue ROI was used (r=0.70, p<0.001). Whole-brain regression results for models that included both scanner and lesion volume as covariates are shown in Supplementary Figure 3A. 0.56, p<0.001), naming (partial r=0.51, p<0.001), and SD performance (partial r=0.46, p=0.004) when only scanner was partialled out. Activity in the whole CSN remained negatively but non-significantly correlated with total lesion volume when scanner was partialled out (r=-0.17, p=0.28). Activity in left and right hemispheric portions of the CSN remained significantly correlated when only scanner was partialled out (partial r=0.79, p<0.001). The same was true when the mirror ROI was used (partial r=0.80, p<0.001), and when the out-of-network homologue ROI was used (r=0.69, p<0.001). Left CSN (partial r = -0.14, p=0.38) and right CSN (partial r=-0.22, p=0.17) activation remained negatively, but less significantly, correlated with lesion volume when scanner was partialled out. Whole-brain regression results for models that included only scanner as a covariate are shown in Supplementary Figure 3B. No covariates: Correlational results are provided in the main text. Whole-brain regression results for models that did not include either scanner or lesion volume as covariates are shown in Figure 3C. Scanner covariate only: Mean CSN activation still predicted fluency (partial r = While our main analyses were performed using lesion volume as a covariate, We note that results were highly similar for all analyses, indicating that scanner and lesion volume covariates had little effect on the results. While the minor effects of scanner are perhaps not surprising due to the use of identical scan parameters, our results suggest that lesion volume differences had only minor effects on functional MRI results in this sample. Supplementary Figure 3. Group analyses with and without different covariates included in the models. A. Whole-brain regressions with both scanner and lesion volume included as covariates. B. Whole-brain and right-hemispheric regressions with scanner included as a covariate. C. Whole-brain regressions with no covariates. Note: RH lesion volume effects are only shown for (B) because lesion volume is the variable of interest in this analysis, and thus scanner is the only applicable covariate. Statistical maps are thesholded at a voxelwise p<0.01 and cluster-corrected at p<0.01 (126 voxels). Importantly, nearly identical results were obtained with and without each covariate included in the models. Supplementary Analysis 4. Semantic decision activation peaks for voxel-wise (FDR) correction thresholds. Supplementary Figure 4. Voxel-wise false discovery rate (vFDR) thresholded semantic decision activation maps for controls (A, vFDR<0.0005) and patients (B, vFDR<0.05) are shown along with tables detailing cluster and peak statistics. These data are intended to (1) demonstrate that the overall patterns of activation are robust even at stringent voxel-wise thresholds, and (2) provide more precise information about the peak locations for each group than the more liberally thresholded maps utilized in the primary analyses. Supplementary Analysis 5. Assessment of damage to "core conceptual network" that showed positive correlations between activity and all three language measures. To assess whether or not differences in activation of the core conceptual network that showed positive relationships to all three language measures might be driven by frequent damage to the network, we assessed lesion frequencies for each voxel included in the network. In addition, we assessed the relationship between mean activation within the network and total damage to the network. The results of these additional analyses are shown in Supplementary Figure 5. Only the left angular gyrus was lesioned in greater than 4 (~10% of) patients, and the voxel with the highest lesion frequency was damaged in 25/43 patients. Activation of the network was weakly negatively (not significantly) related to the amount of damage sustained by the network. Supplementary Figure 5. A. Lesion frequencies for regions where activation showed significant positive relationships to all language measures (i.e. network shown in Figure 3B of the main text) are shown on a partially inflated brain (top), and on slices of a template brain with cross-hairs centered on the voxel with maximum in-mask lesion frequency. Colorbar values indicate the number of patients lesioned at each voxel. Only the left angular gyrus was lesioned in at least 10% (4/43) of patients, with slightly greater than half of patients (25/43) having damage to the left cerebral white matter underlying this region. B. Mean activation within the network showed a negative but non-significant relationship to the extent of damage to the network. This suggests that low levels of activity within this network are not likely to reflect widespread damage to these regions, but rather may reflect the effects of focal damage to regions that were more frequently damaged (such as the left AG) or to inter-regional connections. References for results from literature search shown in Supplementary Table 1. Abel S, Weiller C, Huber W, Willmes K (2014): Neural underpinnings for model- oriented therapy of aphasic word production. Neuropsychologia 57:154–165. Abel S, Weiller C, Huber W, Willmes K, Specht K (2015): Therapy-induced brain reorganization patterns in aphasia. Brain 138:1097–112. Abutalebi J, Rosa PA Della, Tettamanti M, Green DW, Cappa SF (2009): Bilingual aphasia and language control: A follow-up fMRI and intrinsic connectivity study. Brain Lang 109:141–156. Allendorfer J, Kissela B, Holland S, Szaflarski J (2012): Different patterns of language activation in post-stroke aphasia are detected by overt and covert versions of the verb generation fMRI task. Med Sci Monit Monit 18. Belin P, Van Eeckhout P, Zilbovicius M, Remy P, François C, Guillaume S, Chain F, Rancurel G, Samson Y (1996): Recovery from nonfluent aphasia after melodic intonation therapy: a PET study. Neurology 47:1504–1511. role of the right pars opercularis. Ann Neurol 54:310–320. retrieval learning modulates right frontal cortex in patients with left frontal damage. Neuron 36:159–170. Cardebat D (2005): Subcortical aphasia: A longitudinal PET study. Stroke 36:1467– 1473. oxygen level dependent (BOLD) signal in patients with stroke-induced and primary progressive aphasia. NeuroImage Clin 8:87–94. Bonakdarpour B, Parrish TB, Thompson CK (2007): Hemodynamic response function in patients with stroke-induced aphasia: implications for fMRI data analysis. Neuroimage 36:322–31. Behavioral and Neurophysiologic Response to Therapy for Chronic Aphasia. Arch Phys Med Rehabil 90:2026–2033. Brownsett SLE, Warren JE, Geranmayeh F, Woodhead Z, Leech R, Wise RJS (2014): Blank SC, Bird H, Turkheimer F, Wise RJS (2003): Speech production after stroke: The Blasi V, Young AC, Tansy AP, Petersen SE, Snyder AZ, Corbetta M (2002): Word Breier JI, Juranek J, Maher LM, Schmadeke S, Men D, Papanicolaou AC (2009): De Boissezon X, Démonet JF, Puel M, Marie N, Raboyeau G, Albucher JF, Chollet F, Bonakdarpour B, Beeson PM, DeMarco AT, Rapcsak SZ (2015): Variability in blood Eaton KP, Szaflarski JP, Altaye M, Ball AL, Kissela BM, Banks C, Holland SK (2008): Fernandez B, Cardebat D, Demonet JF, Joseph PA, Mazaux JM, Barat M, Allard M Cappa SF, Perani D, Grassi F, Bressi F, Alberoni M, Franceschi M, Bettinardi M, Todde Cognitive control and its impact on recovery from aphasic stroke. Brain 137:242– 54. M, Fazio M (1997): A PET Follow-up Study of Recovery after Stroke in Acute Aphasics. Brain Lang 56:55–67. Cardebat D, Demonet JF, De Boissezon X, Marie N, Mari RM, Lambert J, Baron JC, Puel M (2003): Behavioral and Neurofunctional Changes over Time in Healthy and Aphasic Subjects: A PET Language Activation Study. Stroke 34:2900–2906. Connor LT, DeShazo Braby T, Snyder AZ, Lewis C, Blasi V, Corbetta M (2006): Cerebellar activity switches hemispheres with cerebral recovery in aphasia. Neuropsychologia 44:171–7. Crinion JT, Warburton EA, Lambon-Ralph MA, Howard D, Wise RJS (2006): Listening to narrative speech after aphasic stroke: The role of the left anterior temporal lobe. Cereb Cortex 16:1116–1125. Crinion J, Price CJ (2005): Right anterior superior temporal activation predicts auditory sentence comprehension following aphasic stroke. Brain 128:2858–71. Crosson B, Moore AB, Gopinath K, White KD, Wierenga CE, Gaiefsky ME, Fabrizio KS, Peck KK, Soltysik D, Milsted C, Briggs RW, Conway TW, Gonzalez Rothi LJ (2005): Role of the right and left hemispheres in recovery of function during treatment of intention in aphasia. J Cogn Neurosci 17:392–406. Reliability of fMRI for studies of language in post-stroke aphasia subjects. Neuroimage 41:311–22. (2004): Functional MRI follow-up study of language processes in healthy subjects and during recovery in a case of aphasia. Stroke 35:2171–2176. Fridriksson J (2010): Preservation and modulation of specific left hemisphere regions is vital for treated recovery from anomia in stroke. J Neurosci 30:11558–11564. Fridriksson J, Baker JM, Moser D (2009): Cortical mapping of naming errors in aphasia. Hum Brain Mapp 30:2487–2498. Fridriksson J, Bonilha L, Baker JM, Moser D, Rorden C (2010): Activity in preserved left hemisphere regions predicts anomia severity in aphasia. Cereb Cortex 20:1013–9. Fridriksson J, Morrow-Odom L, Moser D, Fridriksson A, Baylis G (2006): Neural recruitment associated with anomia treatment in aphasia. Neuroimage 32:1403– 1412. GC, Rorden C (2007): Neural correlates of phonological and semantic-based anomia treatment in aphasia. Neuropsychologia 45:1812–22. aphasia recovery. Neuroimage 60:854–63. production after left hemisphere stroke. Neurology 86:1296–1305. following Intermittent Theta Burst Stimulation in Chronic Poststroke Aphasia. Neural Plast 2016:20–23. Fridriksson J, Richardson JD, Fillmore P, Cai B (2012): Left hemisphere plasticity and Geranmayeh F, Leech R, Wise RJS (2016): Network dysfunction predicts speech Griffis JC, Nenert R, Allendorfer JB, Szaflarski JP (2016): Interhemispheric Plasticity Fridriksson J, Moser D, Bonilha L, Morrow-Odom KL, Shaw H, Fridriksson A, Baylis van Hees S, McMahon K, Angwin A, de Zubicaray G, Read S, Copland DA (2014): A Kiran S, Meier EL, Kapse KJ, Glynn PA. (2015): Changes in task-based effective Kurland J, Naeser MA, Baker EH, Doron K, Martin PI, Seekins HE, Bogdan A, Renshaw Heiss WD, Karbe H, Weber-Luxenburger G, Herholz K, Kessler J, Pietrzyk U, Pawlik G functional MRI study of the relationship between naming treatment outcomes and resting state functional connectivity in post-stroke aphasia. Hum Brain Mapp 35:3919–31. Heiss WD, Kessler J, Thiel A, Ghaemi M, Karbe H (1999): Differential capacity of left and right hemispheric areas for compensation of poststroke aphasia. Ann Neurol 45:430–438. (1997): Speech-induced cerebral metabolic activation reflects recovery from aphasia. J Neurol Sci 145:213–217. Jarso S, Li M, Faria A, Davis C, Leigh R, Sebastian R, Tsapkini K, Mori S, Hillis AE (2014): Distinct mechanisms and timing of language recovery after stroke. Cogn Neuropsychol 0:1–22. Karbe H, Thiel A, Weber-luxenburger G, Kessler J, Heiss W (1998): Brain Plasticity in Poststroke Aphasia  : What Is the Contribution of the Right Hemisphere? Brain Lang 230:215–230. connectivity in language networks following rehabilitation in post-stroke patients with aphasia. Front Hum Neurosci 9:1–20. P, Yurgelun-Todd D (2004): Test-retest reliability of fMRI during nonverbal semantic decisions in moderate-severe nonfluent aphasia patients. Behav Neurol 15:87–97. Léger A, Démonet JF, Ruff S, Aithamon B, Touyeras B, Puel M, Boulanouar K, Cardebat D (2002): Neural substrates of spoken language rehabilitation in an aphasic patient: an fMRI study. Neuroimage 17:174–183. Martin PI, Naeser MA, Doron KW, Bogdan A, Baker EH, Kurland J, Renshaw P, Yurgelun-Todd D (2005): Overt naming in aphasia studied with a functional MRI hemodynamic delay design. Neuroimage 28:194–204. Martin PI, Naeser MA, Ho M, Doron KW, Kurland J, Kaplan J, Wang Y, Nicholas M, Baker EH, Fregni F, Pascual-Leone A (2009): Overt naming fMRI pre- and post- TMS: Two nonfluent aphasia patients, with and without improved naming post- TMS. Brain Lang 111:20–35. Mattioli F, Ambrosi C, Mascaro L, Scarpazza C, Pasquali P, Frugoni M, Magoni M, Biagi L, Gasparotti R (2014): Early aphasia rehabilitation is associated with functional reactivation of the left inferior frontal gyrus: a pilot study. Stroke 45:545– 52. Meinzer M, Obleser J, Flaisch T, Eulitz C, Rockstroh B (2007): Recovery from aphasia as a function of language therapy in an early bilingual patient demonstrated by fMRI. Neuropsychologia 45:1247–1256. Meinzer M, Flaisch T, Breitenstein C, Wienbruch C, Elbert T, Rockstroh B (2008): Functional re-recruitment of dysfunctional brain areas predicts language recovery in chronic aphasia. Neuroimage 39:2038–46. Meinzer M, Flaisch T, Obleser J, Assadollahi R, Djundja D, Barthel G, Rockstroh B (2006): Brain regions essential for improved lexical access in an aged aphasic patient: a case report. BMC Neurol 6:28. Meltzer JA, Wagage S, Ryder J, Solomon B, Braun AR (2013): Adaptive significance of Miura K, Nakamura Y, Miura F, Yamada I, Takahashi M, Yoshikawa A, Mizobata T right hemisphere activation in aphasic language comprehension. Neuropsychologia 51:1248–1259. (1999): Functional magnetic resonance imaging to word generation task in a patient with Broca's aphasia. J Neurol 246:939–942. Mohr B, Difrancesco S, Harrington K, Evans S, Pulvermuller F (2014): Changes of right- hemispheric activation after constraint-induced, intensive language action therapy in chronic aphasia: fMRI evidence from auditory semantic processing1. Front Hum Neurosci 8:1–15. Musso M, Weiller C, Kiebel S, Müller SP, Bülau P, Rijntjes M (1999): Training-induced brain plasticity in aphasia. Brain 122 ( Pt 9:1781–90. Naeser MA, Martin PI, Baker EH, Hodge SM, Sczerzenie SE, Nicholas M, Palumbo CL, Goodglass H, Wingfield A, Samaraweera R, Harris G, Baird A, Renshaw P, Yurgelun-Todd D (2004): Overt propositional speech in chronic nonfluent aphasia studied with the dynamic susceptibility contrast fMRI method. Neuroimage 22:29– 41. van Oers CA, Vink M, van Zandvoort MJE, van der Worp HB, de Haan EHF, Kappelle LJ, Ramsey NF, Dijkhuizen RM (2010): Contribution of the left and right inferior frontal gyrus in recovery from aphasia. A functional MRI study in stroke patients with preserved hemodynamic responsiveness. Neuroimage 49:885–93. Papoutsi M, Stamatakis EA, Griffiths J, Marslen-Wilson WD, Tyler LK (2011): Is left fronto-temporal connectivity essential for syntax? Effective connectivity, tractography and performance in left-hemisphere damaged patients. Neuroimage 58:656–64. Peck KK, Moore AB, Crosson BA, Gaiefsky M, Gopinath KS, White K, Briggs RW (2004): Functional Magnetic Resonance Imaging before and after Aphasia Therapy: Shifts in Hemodynamic Time to Peak during an Overt Language Task. Stroke 35:554–559. Perani D, Cappa SF, Tettamanti M, Rosa M, Scifo P, Miozzo A, Basso A, Fazio F (2003): A fMRI study of word retrieval in aphasia. Brain Lang 85:357–368. Postman-Caucheteux WA, Birn RM, Pursley RH, Butman JA, Solomon JM, Picchioni D, McArdle J, Braun AR (2010): Single-trial fMRI shows contralesional activity linked to overt naming errors in chronic aphasic patients. J Cogn Neurosci 22:1299–1318. reorganization of language in both hemispheres of patients with chronic aphasia. Neuroimage 28:481–9. Raboyeau G, De Boissezon X, Marie N, Balduyck S, Puel M, Bézy C, Démonet JF, Cardebat D (2008): Right hemisphere activation in recovery from aphasia: Lesion effect or function recruitment? Neurology 70:290–298. Richter M, Miltner WHR, Straube T (2008): Association between therapy outcome and right-hemispheric activation in chronic aphasia. Brain 131:1391–1401. Robson H, Zahn R, Keidel JL, Binney RJ, Sage K, Lambon Ralph MA (2014): The anterior temporal lobes support residual comprehension in Wernicke's aphasia. Brain 137:931–43. Rochon E, Leonard C, Burianova H, Laird L, Soros P, Graham S, Grady C (2010): Neural changes after phonological treatment for anomia: An fMRI study. Brain Pulvermüller F, Hauk O, Zohsel K, Neininger B, Mohr B (2005): Therapy-related Rosen HJ, Petersen SE, Linenweber MR, Snyder AZ, White DA, Chapman L, Seghier ML, Bagdasaryan J, Jung DE, Price CJ (2014): The Importance of Premotor Specht K, Zahn R, Willmes K, Weis S, Holtel C, Krause BJ, Herzog H, Huber W (2009): Lang 114:164–179. Dromerick a W, Fiez JA, Corbetta MD (2000): Neural correlates of recovery from aphasia after damage to left inferior frontal cortex. Neurology 55:1883–1894. Sakatani K, Xie Y, Wemara L, Li S, Zuo H (1998): Language-Activated Cerebral Blood Oxygenation and Hemodynamic Changes of the Left Prefrontal Cortex in Poststroke Aphasic Patients. Stroke 29:12–14. Saur D, Lange R, Baumgaertner A, Schraknepper V, Willmes K, Rijntjes M, Weiller C (2006): Dynamics of language reorganization after stroke. Brain 129:1371–84. Saur D, Ronneberger O, Kümmerer D, Mader I, Weiller C, Klöppel S (2010): Early functional magnetic resonance imaging activations predict language outcome after stroke. Brain 133:1252–1264. Cortex for Supporting Speech Production after Left Capsular-Putaminal Damage. J Neurosci 34:14338–14348. Sharp DJ, Turkheimer FE, Bose SK, Scott SK, Wise RJS (2010): Increased frontoparietal integration after stroke and cognitive recovery. Ann Neurol 68:753–6. Sims JA, Kapse K, Glynn P, Sandberg C, Tripodis Y, Kiran S (2016): The relationships between the amount of spared tissue, percent signal change, and accuracy in semantic processing in aphasia. Neuropsychologia 84:113–126. Joint independent component analysis of structural and functional images reveals complex patterns of functional reorganisation in stroke aphasia. Neuroimage 47:2057–2063. Spironelli C, Angrilli A (2015): Brain plasticity in aphasic patients: intra- and inter- hemispheric reorganisation of the whole linguistic network probed by N150 and N350 components. Sci Rep 5:12541. Szaflarski JP, Eaton K, Ball AL, Banks C, Vannest J, Allendorfer JB, Page S, Holland SK (2011a): Poststroke aphasia recovery assessed with functional magnetic resonance imaging and a picture identification task. J Stroke Cerebrovasc Dis 20:336–345. Szaflarski JP, Allendorfer JB, Banks C, Vannest J, Holland SK (2013): Recovered vs. not-recovered from post-stroke aphasia: the contributions from the dominant and non-dominant hemispheres. Restor Neurol Neurosci 31:347–60. Szaflarski JP, Vannest J, Wu SW, DiFrancesco MW, Banks C, Gilbert DL (2011b): Excitatory repetitive transcranial magnetic stimulation induces improvements in chronic post-stroke aphasia. Med Sci Monit 17:CR132-9. Thiel A, Hartmann A, Rubi-Fessen I, Anglade C, Kracht L, Weiduschat N, Kessler J, Rommel T, Heiss W-D (2013): Effects of noninvasive brain stimulation on language networks and recovery in early poststroke aphasia. Stroke 44:2240–6. processing in aphasia: Changes in lateralization patterns during recovery reflect cerebral plasticity in adults. Electroencephalogr Clin Neurophysiol 102:86–97. Thompson CK, Riley EA, den Ouden DB, Meltzer-Asscher A, Lukic S (2013): Training verb argument structure production in agrammatic aphasia: Behavioral and neural recovery patterns. Cortex 49:2358–2376. Thomas C, Altenmuller E, Marckmann G, Kahrs J, Dichgans J (1997): Language Vitali P, Abutalebi J, Tettamanti M, Danna M, Ansaldo A-I, Perani D, Joanette Y, Cappa Thompson CK, den Ouden DB, Bonakdarpour B, Garibaldi K, Parrish TB (2010): Neural Thompson CK, Bonakdarpour B, Fix SF (2010): Neural mechanisms of verb argument structure processing in agrammatic aphasic and healthy age-matched listeners. J Cogn Neurosci 22:1993–2011. plasticity and treatment-induced recovery of sentence processing in agrammatism. Neuropsychologia 48:3211–3227. Tyler LK, Marslen-Wilson WD, Randall B, Wright P, Devereux BJ, Zhuang J, Papoutsi M, Stamatakis EA. (2011): Left inferior frontal cortex and syntax: Function, structure and behaviour in patients with left hemisphere damage. Brain 134:415– 431. SF (2007): Training-induced brain remapping in chronic aphasia: a pilot study. Neurorehabil Neural Repair 21:152–60. Warburton E, Price CJ, Swinburn K, Wise RJ (1999): Mechanisms of recovery from aphasia: evidence from positron emission tomography studies. J Neurol Neurosurg Psychiatry 66:155–61. Warren JE, Crinion JT, Lambon Ralph MA, Wise RJS (2009): Anterior temporal lobe connectivity correlates with functional outcome after aphasic stroke. Brain 132:3428–42. Weiduschat N, Thiel A, Rubi-Fessen I, Hartmann A, Kessler J, Merl P, Kracht L, Rommel T, Heiss WD (2011): Effects of repetitive transcranial magnetic stimulation in aphasic stroke: a randomized controlled pilot study. Stroke 42:409–15. Noth J, Diener HC (1995): Recovery from Wernicke's aphasia: A positron emission tomographic study. Ann Neurol 37:723–732. Winhuisen L, Thiel A, Schumacher B, Kessler J, Rudolf J, Haupt WF, Heiss WD (2005): Role of the contralateral inferior frontal gyrus in recovery of language function in poststroke aphasia: a combined repetitive transcranial magnetic stimulation and positron emission tomography study. Stroke 36:1759–63. The right inferior frontal gyrus and poststroke aphasia: a follow-up investigation. Stroke 38:1286–92. Zahn R, Drews E, Specht K, Kemeny S, Reith W, Willmes K, Schwarz M, Huber W (2004): Recovery of semantic word processing in global aphasia: A functional MRI study. Cogn Brain Res 18:322–336. Zhu D, Chang J, Freeman S, Tan Z, Xiao J, Gao Y, Kong J (2014): Changes of functional connectivity in the left frontoparietal network following aphasic stroke. Front Behav Neurosci 8:167. Weiller C, Isensee C, Rijntjes M, Huber W, Müller S, Bier D, Dutschka K, Woods RP, Winhuisen L, Thiel A, Schumacher B, Kessler J, Rudolf J, Haupt WF, Heiss WD (2007):
1702.07227
1
1702
2017-02-20T10:33:03
The knowledge paradox: why knowing more is knowing less
[ "q-bio.NC" ]
To provide an explanation of the evolution of scientific knowledge, I start from the assumption that knowledge is based on concepts, and propose that each concept about reality is affected by vagueness. This entails a paradox, which I term Knowledge Paradox (KP): i.e. we need concepts to acquire knowledge about the real world but each concept is a step away from reality. The KP provides a unifying context for the sorites and the liar paradoxes. Any concept is viewed as a sorites, i.e. it is impossible to set a boundary between what is, and what is not, the entity to which the concept refers. Hence, any statement about reality can be reduced to a liar, wherefrom the KP follows in its most general form: -If I know, then I do not know-. The KP is self-referential but not contradictory, as it can be referred to two levels of knowledge: -if I know(epistemic), then I do not know(ontic)-, where the ontic level is made unachievable by concept vagueness. Such an interpretation of scientific knowledge provides an understanding of its dynamics. Concept proliferation within theories produces periods of knowledge decay that are episodically reversed by the formulation of new theories based on a smaller number of synthetic concepts.
q-bio.NC
q-bio
The knowledge paradox: why knowing more is knowing less Bruno Burlando Dipartimento di Sienze e Innovazione Tecnologica Università del Piemonte Orientale Viale Teresa Michel 11 15121 Alessandria Italy e-mail : [email protected] 1 Abstract To provide an explanation of the evolution of scientific knowledge, I start from the assumption that knowledge is based on concepts, and propose that each concept about reality is affected by vagueness. This entails a paradox, which I term Knowledge Paradox (KP): i.e. we need concepts to acquire knowledge about the real world but each concept is a step away from reality. The KP provides a unifying context for the sorites and the liar paradoxes. Any concept is viewed as a sorites, i.e. it is impossible to set a boundary between what is, and what is not, the entity to which the concept refers. Hence, any statement about reality can be reduced to a liar, wherefrom the KP follows in its most general form: "If I know, then I do not know". The KP is self-referential but not contradictory, as it can be referred to two levels of knowledge: "if I knowepistemic, then I do not knowontic", where the ontic level is made unachievable by concept vagueness. Such an interpretation of scientific knowledge provides an understanding of its dynamics. Concept proliferation within theories produces periods of knowledge decay that are episodically reversed by the formulation of new theories based on a smaller number of synthetic concepts. . 2 Introduction This article concerns an analysis of the meaning and progress of scientific knowledge. I will not make attempts at defining the limits of science, i.e. dealing with the long debated problem of demarcation (Resnik, 2000). Scientific knowledge will be simply defined here as a particularly structured form of conceptual knowledge, deriving from perceptual or instrumental experience of the world. It is generally assumed that the purpose of science is to develop laws and theories in order to explain, understand and predict phenomena (Chalmers, 1990). Accordingly, it is believed that the scientific method has been one of the driving forces of human progress, while our knowledge about the world is usually related to the ability of theories to make predictions about reality (Gauch, 2003). However, progressive accumulation of scientific evidence can reveal shortcomings in extant paradigms, ultimately leading to theoretical rearrangements. Such a trend has been emphasized by Kuhn, with his distinction between normal and revolutionary phases in the progress of science (Kuhn 1962). Cognitive limits of concepts I start from the assumption that the failure of scientific theories derives from intrinsic cognitive limits of concepts. The term 'concept' will be used here in a wide acception, i.e. as an idea or notion that serves to designate a category of entities, or more in general, to provide a model of reality. Even though non-conceptual content of thought has been considered in the analysis of human cognitive processes and awareness (Bermudez, 1995), it is generally believed that conceptual content of thought is typical of scientific reasoning (Achinstein, 1968). Hence, our mind uses concepts in order to describe the real world, and such a process is generally considered an essential step in the advancement of science. I wish to stress that the shortcomings of scientific paradigms, which may lead to a collapse of theoretical buildings, would depend on the reduction of the world to a set of concepts, i.e. discrete 3 model objects, thereby yielding an artifactual vision of reality. A vivid reminiscence of these ideas can be found in the words of the French poet and philosopher Paul Valery: «Tout ce qui est simple est faux, mais tout ce qui ne l'est pas est inutilisable» (Guégan and Renaud, 2004). On a similar ground, my analysis holds that concepts can be viewed as an unavoidable simplification of reality, which inevitably leads to its falsification. These arguments are to some extent related to instrumentalism (White, 1943), which assumes that concepts and theories are merely useful instruments, or 'fictions', whose function is not to depict reality but rather to provide descriptions and predictions about the physical world. However, I am pushing my analysis further on, by proposing that concepts introduce elements of inconsistency within scientific theories that sooner or later lead to thorough revisions. Concept vagueness and the sorites The failure of concepts to provide a reliable description of the world addresses the problem of vagueness. This latter is strictly related to the sorites paradox, which entails the possibility that the borderline between a predicate and its negation be fuzzy (Tye, 1994; Keefe, 2000). In his analysis of the sorites, James Cargile refers to the impossibility of setting an exact point of transition between a developing tadpole and an adult frog (Cargile, 1969). It may be assumed that a proper example of vagueness had to be chosen, whereas in contrast my argument is aimed at showing that any real entity would have been appropriate, since I assume that vagueness is intrinsic to concepts and, therefore, any concept is susceptible to the sorites. Obviously, the vagueness of concepts does not affect the practical use of the real objects to which they refer. In general, vagueness is ruled out whenever a certain approximation of reality can be tolerated in order to accomplish a particular project, as commonly occurs in everyday life, technical fields and professional work. Conversely, the vagueness of concepts heavily affects the rigorous in-depth analyses that need to be deployed in scientific activities. 4 Fundamental concepts of physics, like matter, time, and space, providing basic reference frames to scientific theories, are affected by vagueness. After the arising of quantum mechanics, there has been growing awareness that space and time are vague on a microscopic scale (Dolan, 2005). Moreover, their vagueness does not disappear on a macroscopic scale, as shown by the problem of infinite divisibility highlighted by Zeno's paradoxes (Mueller 1969), while their unwarranted use can lead to the puzzling situations of time travel paradoxes (Nahin, 1999). Also, the concept of matter has been made uncertain by quantum physics, under which the identity of elementary particles is suggested to be irreducibly vague (French and Krause, 2003; Chibeni, 2004). In the field of biology, the formulation of the concept of cell has been a milestone of enormous theoretical importance (Baker, 1953). However, the identity of cells within organisms remains vague, due to the occurrence of syncytial tissues, endocytobiosis, incomplete cell divisions, and exchange of material between a cell and its surroundings. Similar reasoning can be applied to a number of biological entities. In summary, science is affected by vagueness at its deepest foundations. The formulation of so- called exact laws is only possible if an appropriate armamentarium of concepts allows the edification of theoretical systems within which these laws are meaningful, as occurs e.g. for gravitation, electromagnetism and thermodynamics. By contrast, the lack of a suitable set of concepts renders extremely difficult the determination of exact laws for nonlinear systems, such as fluid turbulence, living beings, and stock markets (Bar-Yam, 1997). Knowledge Paradox and the progress of science Assuming that any concept about reality generates a sorites, since it involves some discretization of the real world, it follows that the formulation of concepts gives rise to vagueness and move us away from reality. This configures a paradox that I call the Knowledge Paradox (KP): scientific activities yield discoveries that are implemented within extant theories by means of new, ad hoc concepts, thus producing increasingly complex conceptual frameworks that progressively put our 5 vision of reality out of focus. Consequently, the accumulation of concepts within theories does not produce an advancement of knowledge, but rather a progressive decay. In order to clarify such a counterintuitive idea, I will use an example from the field of molecular biology. If a new, bona-fide protein is found, say protein A, a new concept is needed to implement such a discovery within the molecular theory of life. However, it is not clear if the notion of protein A refers to the protein's native form, to one of its post-translational modifications, or to one of the interactions established by the protein with other molecules, including the bulk of solvent water molecules. Now, given that each molecule of protein A is in a particular chemico-physical status as regards to its interactions with other molecules, a specific concept would be needed for each protein A molecule. Yet, even if this impracticable conceptualization could be done, it would not solve the problem, as molecular interactions are subjected to continuous change along time. Hence, the concept of the protein A is vague, representing at most an approximation of reality, and consequently it adds vagueness to the molecular theory of life. As stated above, this does not imply that the new concept cannot be helpful at all, but its use will be only profitable within theoretical boundaries entailing a simplification of reality. Conversely, if the aim is to achieve a better understanding of reality, the new concept will eventually become a hindrance. As shown above, the accumulation of concepts negatively affects the description of reality. However, from time to time exhaustive revisions of data may succeed in producing synthetic theories that explain facts by a smaller number of new, revolutionary concepts. The evolution of scientific knowledge would therefore consist of episodic advancements, characterized by a reduction in the number of concepts, interspersed among long periods of decay characterized by concept accumulation (Fig. 1). Such a pattern can be easily recognized. By introducing his theory of natural selection, Darwin (1859) provided a synthetic explanation for the astonishing set of life diversity that had been collected through the years. About one century later, the theory of the genetic code yielded a unifying conceptual basis for biological heredity (Crick, 1967). Wegener's continental drift (1915), and the ensuing theory of plate tectonics, provided a synthetic explanation 6 for the movement of Earth's plates, earthquakes, volcanoes, oceanic trenches, and mountain range formation. Also, by introducing the concept of "quark" in the 1960s, Gell-Mann and Zweig offered a unifying account of the variety of hadron particles described during the first half of the 20th century (Lichtenberg and Rosen, 1980). Similar features are shared by other major scientific breakthroughs, such as those carried out by Newton, Maxwell and Freud (Kuhn, 1962). Knowledge Paradox and the liar If no theory can describe the real world in a completely affordable way, this must hold true for the KP itself. In other words, if any concept is vague when applied to reality, then the concept of 'concept' is vague too, and consequently also the KP is affected by vagueness. Hence, the KP makes a statement that, in the ultimate analysis, concerns its own meaning, and therefore is equivalent to the liar paradox (Martin, 1978). The liar is endowed with self-referentiality, or in other terms, it implies and is implied by its negation (Suber, 1990). Hence, it represents a contradiction built into the structure of language, while on the other hand the KP entails an endless loop within the structure of knowledge. Similar to the antinomy of the liar: "If this sentence is true, then it is not true", the KP can be formulated as: "If I know, then I do not know", which seems to make contradiction unavoidable. A solution to the liar has been proposed by Tarski (1956) who, in his approach to the semantic concept of truth, adopted a hierarchy of languages consisting of object-language, to which the definition of truth is applied, and of meta-language, providing the definition of truth for object- language. By following a similar approach, knowledge can be analyzed from the standpoint of a meta-knowledge level, by making a distinction between our ideas about physical entities, viz. the epistemic level, and the real status of these entities, viz. the ontic level (Atmanspacher & Primas, 2003). Such a distinction allows a Tarskian reformulation of the KP providing two meanings for 'I know', whereby it should read: "If I knowepistemic, then I do not knowontic". 7 According to this view, the formulation of concepts generates a gap between ontic and epistemic knowledge that cannot be amended, since any attempt at reducing this gap can only be carried out by formulating other concepts, and so on. Hence, the KP sets the limits of knowledge, while its self- referentiality highlights the unamendability of these limits. A symptom of the ontic/epistemic disjunction of knowledge is the arising of inconsistencies in scientific theories that prelude the advent of their crisis. In Physics, striking examples have been the discoveries of the photoelectric effect and the black body radiation, which raised inconsistencies in classic physics and opened the way to quantum mechanics (Planck, 1901). In Biology, the Central Dogma of molecular genetics holds that biological information flows from DNA or RNA to proteins, and not the reverse (Crick, 1958). However, the discovery of prions (Prusiner, 1982), and the finding in yeast that these proteins can transmit information through generations and modify the expression of the genome (True & Lindquist, 2000), has posed a serious challenge to the overall consistency of the Dogma. Inconsistency of knowledge Combined analyses of the liar and the sorites have been carried out with the aim of finding a unifying treatment (Tappenden, 1993). The KP provides a link between the sorites and the liar by showing that any statement about the world is, or is entailed by, a statement that makes use of a sorites, and can therefore be reconducted to a liar. In the above example of protein A, the sorites emerges from the impossibility of setting a boundary between what is protein A and what is not protein A. Hence, the liar follows: "If this entity is (epistemologically) protein A, then it is not (ontologically) protein A". In summary, knowledge is intrinsically inconsistent. So far as we formulate new concepts, then, the more concepts we have, the more inconsistencies we raise, the more we worsen knowledge about the world. By contrast, if we manage to reduce the number of concepts, by formulating synthetic ones, we reduce inconsistencies and eventually obtain an improvement of knowledge. So, 8 contrarily to what it may be intuitively assumed: the more concepts, the less knowledge, and vice versa. References Achinstein, P. 1968. Concepts of Science: A Philosophical Analysis. Baltimore: The Johns Hopkins University Press. Atmanspacher, H., Primas, H., 2003. Epistemic and ontic quantum realities. In: Castell L., Ischebeck O. (Eds), Time, Quantum and Information, Berlin: Springer, pp 301–321. Bar-Yam, Y. 1997. Dynamics of Complex Systems. Cambridge, MA: Perseus Books. Baker, J.R. 1953. The cell-theory: a restatement, history and critique. Quarterly Journal of Microscopical Science 94: 407-440. Bermudez, J. L. 1995. Non-conceptual content: From perceptual experience to subpersonal computational states. Mind and Language 10: 333-369. Cargile, J. 1969. The sorites paradox. Br. J. Phyl. Sci. 20: 193-202. Chalmers, A. 1990. Science and its Fabrication. Minneapolis, MN: University of Minnesota Press. Chibeni, S.S. 2004. Ontic vagueness in microphysics. Sorites 15: 29-41. Retrieved February 14, 2017, from http://www.sorites.org/Issue_15/chibeni.htm Crick, F.H.C. 1967. The Croonian lecture, 1966. The genetic code. Proc. R. Soc. Lond. Ser. B 167: 331-347. Crick, F.H.C. 1958. Central dogma of molecular biology. Nature 227: 561- 563. Darwin, C. 1859. On the Origin of Species. London: John Murray. Dolan, R.P. 2005. The inflaton spacetime model: making sense of the standard models of particle physics and cosmology. Retrieved February 15, 2017 from http://home.earthlink.net/~dolascetta/ISTModel.pdf French, S. and Krause, D. 2003. Quantum vagueness. Erkenntnis 59: 97-124. Gauch, H.G.Jr. 2003. Scientific Method in Practice. Cambridge, UK: Cambridge University Press. 9 Guégan, J.F., Renaud, F. 2004. Vers une écologie de la santé. In Barbault R., Chevassus-au-Louis B. (Eds.): Biodiversité et Changements Globaux. ADPF/MAE, Paris, France. pp. 100-135. Keefe, R. 2000. Theories of vagueness. Cambridge, UK: Cambridge University Press. Kuhn, T.S. 1962. The Structure of Scientific Revolutions. Chicago, IL: University of Chicago Press. Lichtenberg, D.B., Rosen, S.P. (Eds) 1980. Developments in the Quark Theory of Hadrons, Vol. 1. 1964-1978. Nonantum, MA: Hadronic Press. Mueller, I. 1969. Zeno's paradoxes and continuity. Mind 78: 129-131. Martin, R. (Ed.) 1978. The Paradox of the Liar. Atascadero, CA: Ridgeview. Nahin, P.J. 1999. Time Machines: Time Travel in Physics, Metaphysics and Science Fiction. New York: Springer-Verlag. Planck, M. 1901. Über das Gesetz der Energieverteilung im Normalspektrum. Annalen Der Physik 4: 553 ff. Prusiner, SB. 1982. Novel proteinaceous infectious particles cause scrapie. Science 216 : 136-144. Resnik, D.B. 2000. A pragmatic approach to the demarcation problem. Studies in History and Philosophy of Science Part A 31: 249-267. Suber, P. 1990. The Paradox of Self-Amendment: A Study of Law, Logic, Omnipotence, and Change. New York: Peter Lang Publishing. Tappenden, J. 1993. The Liar and sorites paradoxes: Toward a unified treatment. The Journal of Philosophy 90: 551-577. Tarski, A. 1956. The Concept of Truth in Formalized Languages', Logic, Semantics and Metamathematics. Oxford: Oxford University Press. True, H.L., Lindquist, S.L. 2000. A yeast prion provides a mechanism for genetic variation and phenotypic diversity. Nature 28: 477-483. Tye, M. 1994. Sorites paradoxes and the semantics of vagueness. In Tomberlin J. (Ed.), Philosophical Perspectives: Logic and Language. Atascadero, CA: Ridgeview. Wegener, A. 1915. Die Entstehung der Kontinente und Ozeane. Vieweg, Braunschweig. 10 White, M. 1943. The Origin of Dewey's Instrumentalism. New York: Columbia University Press. 11 Figures K knowledge decay knowledge decay more concepts needed to implement new data new theory with synthe,c concepts more concepts needed to implement new data t Figure 1. A hypothetical knowledge index (K), varying along time (t), is defined as being inversely proportional to the number of concepts used within a theory. So far as increasing numbers of concepts are needed to implement newly acquired data, the index scales down, indicating a decay of knowledge. However, from time to time a thorough revision of data gives rise to a synthetic theory consisting of a few innovative concepts. This renders many concepts obsolete, thus producing a sudden improvement of scientific knowledge. 12
1111.6493
1
1111
2011-11-22T15:38:02
Depressive patients are more impulsive and inconsistent in intertemporal choice behavior for monetary gain and loss than healthy subjects- an analysis based on Tsallis' statistics
[ "q-bio.NC" ]
Depression has been associated with impaired neural processing of reward and punishment. However, to date, little is known regarding the relationship between depression and intertemporal choice for gain and loss. We compared impulsivity and inconsistency in intertemporal choice for monetary gain and loss (quantified with parameters in the q-exponential discount function based on Tsallis' statistics) between depressive patients and healthy control subjects. This examination is potentially important for advances in neuroeconomics of intertemporal choice, because depression is associated with reduced serotonergic activities in the brain. We observed that depressive patients were more impulsive and time-inconsistent in intertemporal choice action for gain and loss, in comparison to healthy controls. The usefulness of the q-exponential discount function for assessing the impaired decision-making by depressive patients was demonstrated. Furthermore, biophysical mechanisms underlying the altered intertemporal choice by depressive patients are discussed in relation to impaired serotonergic neural systems. Keywords: Depression, Discounting, Neuroeconomics, Impulsivity, Inconsistency, Tsallis' statistics
q-bio.NC
q-bio
1 [ http://node.nel.edu/?node_id=7412 NEL 2008] Depressive patients are more impulsive and inconsistent in intertemporal choice behavior for monetary gain and loss than healthy subjects- an analysis based on Tsallis' statistics. Taiki Takahashi1, Hidemi Oono2, Takeshi Inoue3, Shuken Boku3, Yuki Kako 3, Yuji Kitaichi3, Ichiro Kusumi3, Takuya Masui3, Shin Nakagawa3, Katsuji Suzuki3, Teruaki Tanaka3, Tsukasa Koyama3, and Mark H. B. Radford4 1Direct all correspondence to Taiki Takahashi, Unit of Cognitive and Behavioral Sciences Department of Life Sciences, School of Arts and Sciences, The University of Tokyo, Komaba, Meguro-ku, Tokyo, 153-8902, Japan ([email protected]). 2Department of Behavioral Science, Faculty of Letters, Hokkaido University, N.10, W.7, Kita-ku, Sapporo, 060-0810, Japan 3Department of Psychiatry, Graduate School of Medicine, Hokkaido University, N.15, W.7, Kita-ku, Sapporo, 060-8638 4Symbiosis Group Limited, P.O. Box 1192, Milton, 4064 Australia; and Department of Behavioral Science, Hokkaido University, N.10, W.7, Kita-ku, Sapporo, 060-0810, Japan Acknowledgements: The research reported in this paper was supported by a grant from the Grant- in-Aid for Scientific Research (“21st century center of excellence” grant) from the Ministry of Education, Culture, Sports, Science and Technology of Japan. 2 Abstract Depression has been associated with impaired neural processing of reward and punishment. However, to date, little is known regarding the relationship between depression and intertemporal choice for gain and loss. We compared impulsivity and inconsistency in intertemporal choice for monetary gain and loss (quantified with parameters in the q-exponential discount function based on Tsallis' statistics) between depressive patients and healthy control subjects. This examination is potentially important for advances in neuroeconomics of intertemporal choice, because depression is associated with reduced serotonergic activities in the brain. We observed that depressive patients were more impulsive and time-inconsistent in intertemporal choice action for gain and loss, in comparison to healthy controls. The usefulness of the q-exponential discount function for assessing the impaired decision-making by depressive patients was demonstrated. Furthermore, biophysical mechanisms underlying the altered intertemporal choice by depressive patients are discussed in relation to impaired serotonergic neural systems. Keywords: Depression, Discounting, Neuroeconomics, Impulsivity, Inconsistency, Tsallis' statistics 3 1. Introduction 1.1 Impulsivity and depression Depressive patients have neuropsychological impairments including deficits in sensitivity to reward and punishment [1] and decision-making [2]. It is also known that altered serotonergic neural systems are associated with depression and related decision-making impairments [3]. Because depressive disorders are strongly associated with impulsive and risky behavior such as suicide attempts [4,5], it is of importance to establish a rigorous behavioral framework for assessing the degrees of impairments in decision-making by depressive subjects. In this study, we examined two distinct neuropsychological tendencies; i.e., impulsivity and inconsistency in intertemporal choice (delay discounting) by clinically diagnosed depressive patients. Briefly, impulsivity in intertemporal choice refers to the degree of preference for (or aversion to) smaller sooner rewards (punishments) over later larger ones; while inconsistency in intertemporal choice refers to time-dependency of the impulsivity in intertemporal choice (we will illustrate these two tendencies in the next section). Because previous studies indicate that serotonergic functioning in the brain (known to be reduced in depressed patients) may be related to the evaluation of future rewards [6], these investigations are of potential importance for neuroeconomic understandings of depression and biophysical mechanisms of serotonergic systems underlying impaired decision-making by mood disorder patients. However, to date, little is known about the relationships between depression, impulsivity and inconsistency in intertemporal choice, partly due to a difficulty in operationalizing impulsivity and inconsistency in intertemporal choice in a distinct manner. Recent studies in econophysics and neuroeconomics have demonstrated the usefulness of the q-exponential discount function [7-9] to parametrize both impulsivity and inconsistency in intertemporal choice. We therefore compared both impulsivity and inconsistency in intertemporal choice for monetary gains and losses between depressive patients and healthy control subjects, by utilizing the q-exponential discount function. It is further to be noted that this study is the first to compare both impulsivity and time-inconsistency in intertemporal choice between control and neuropsychiatric patient groups by utilizing the q-exponential discount function based on Tsallis' statistics. 1.2 Impulsivity and inconsistency in intertemporal choice Impulsivity in intertemporal choice (delay discounting) for gain refers to preference for smaller but more immediate rewards over larger but more delayed ones. Consider the following example: 4 (A) Choose between (A.1) One cup of coffee now. (A.2) Two cups of coffee tomorrow. (B) Choose between (B.1) One cup of coffee in one year. (B.2) Two cups of coffee in [one year plus one day]. Most people may prefer a sooner smaller reward in (A) (i.e., (A.1), impulsive choice); while prefer a later larger reward in (B) (i.e., (B.2), patient choice). These examples demonstrate that most people are patient in making a plan about intertemporal choice in the distant future, but impulsive in the near future, resulting in "preference reversal" as time passes [10-12]. It is to be noted that consistent decision-makers should choose either [(A1) and (B1)] or [(A2) and (B2)] because time-intervals between sooner smaller rewards and later larger ones are the same (i.e., one day) in both (A) and (B). In summary, impulsivity in delay discounting of gain corresponds to the degree to which the subject discount the delayed reward; while inconsistency in delay discounting corresponds to the dependency of the intensity of aversion to waiting for one day to obtain an additional cup of coffee, on time-points (now or 1 year later) in the examples above. It is also to be noted that example A is about the actual intertemporal choice action; on the other hand, example B is about the intertemporal choice plan. Note that people cannot actually take future actions now, and therefore choosing (B2) is a future plan (not an actual action); while choosing (A1) is an actual action. Neuroeconomic studies have reported that addiction to drugs of abuse is associated with impulsive intertemporal choice for gain (e.g. choosing immediate rewards from drug intake at the cost of later larger rewards such as healthy body in later life) [13-16]; while little is known about the relationships between neuropsychiatric illnesses including addiction and inconsistency in intertemporal choice. In intertemporal choice for loss, discounting of delayed loss corresponds to a decrease in aversion to loss when the loss is delayed. In other words, subjects who avoid paying small costs immediately and choose to pay larger later costs are strong delay discounters of loss. Therefore, strong delay discounting of loss (impulsivity in delay discounting of loss) represents the marked tendency of procrastination about paying a cost. 1.3 q-exponential discount function based on Tsallis' statistics Recent econophysical and neuroeconomic studies [7-9] proposed and examined the following q-exponential discount function based on Tsallis' statistics: 5 V(D)=A/ expq(kqD)=A/[1+(1-q)kqD] 1/(1-q) (Equation 1) where expq () is the q-exponential function, D is a delay until receipt of a reward, A is the value of a reward at D=0, and kq is a parameter of impulsivity at delay D=0. Note that when q=0, equation 1 is the same as a simple hyperbolic discount function [10-12]: V(D)=A/(1+khD), (Equation 2), where kh=k0, while q→1, is the same as an exponential discount function proposed in classical economics [17]: V(D)=Aexp(-keD), (Equation 3) where ke=k1. In any continuous time-discounting functions, a discount rate is defined as -(dV(D)/dD)/V(D), independently of functional types of discount models, and larger discount rates corresponds to more impulsive intertemporal choice. In the q-exponential discount function, the discount rate (q-exponential discount rate) is: -V’(D)/V(D)=kq/(1+kq(1-q)D). (Equation 4) We can see that when q=1, the discount rate is independent of delay D, corresponding to exponential discounting (consistent intertemporal choice); while for q<1, the discount rate is a decreasing function of delay D, resulting in preference reversal. This can also be demonstrated by calculating the time-derivative of the q-exponential discount rate: (d /dD) [-V’(D)/V(D)]= kq 2(1-q)/(kq(1-q)D+1)2 (Equation 5) which is negative and positive for q<1 and q>1, respectively. Also, impulsivity at delay D=0 is equal to kq irrespective of q. We have previously shown that the q-exponential discount function is capable of continuously quantifying human subjects' inconsistency in intertemporal choice [8,9]. Namely, human agents with smaller q values are more inconsitent in intertemporal choice. If q is less than 0, the intertemporal choice behavior is more inconsistent than hyperbolic discounting. 1.4 Objectives of the present study The aim of this study was to examine impulsivity (kq in Equation 1) and inconsistency (Equation 2) in delay discounting of gain and loss among depressed patients, in comparison to healthy people. Based on previous research that has suggested higher 6 impulsivity among depressed patients, we propose the hypothesis that the depressed patients are more impulsive in intertemporal choice behavior than healthy normal subjects. It is important to note that our present study is the first to compare inconsistency in intertemporal choice between healthy controls and neuropsychiatric patients. Notably, previous findings regarding the relationship between serotonin and discounting have been mixed. A rodent animal model reported that a reduction in serotonergic activities were associated with exaggerated impulsivity in intertemporal choice over several seconds [18]; while a psychopharmacological study with human subjects did not observe a significant effect of a reduction in serotonergic activities on intertemporal choice over a year [19]. How our present study resolves the discrepancy between these studies is also discussed after presentation of our experimental data. Furthermore, we also aimed to examine the differences in impulsivity and inconsistency between discounting delayed gain and loss. The rationale is that no study to date examined the effect of the sign of delayed outcomes (i.e., gain or loss) on inconsistency in delay discounting, although it has been reported that delayed gain is more rapidly discounted than delayed loss (i.e., the sign effect in intertemporal choice [12]). 2. Methods 2.1 Participants Participants were 29 depressive patients diagnosed with DSM-IVTR (major depressive disorder and bipolar disorder, most recent episode: depressed) and 15 healthy control subjects. The depressive patient participants were patients of Hokkaido University Hospital, which serves a large urban catchments area. Participants’ ages ranged from 27 to 67, with a mean age of 43.83 years (SD = 2.25), and healthy controls’ ages ranged from 31 to 71 with a mean age of 47.6 years (SD = 3.53). All participants including depressive patients and healthy control subjects filled an informed consent form before starting the experiment. The group of depressive patients consisted of both unipolar (i.e., patients with only a major depression phase) and bipolar (i.e., patients with both depressive and manic phases) disorders. It is important to note that bipolar disorder patients in a depressed state at the time of their participation (not in a manic state) were included in the present study, in order to exclude the influences of manic mental states on intertemporal choice. Consequently, there was no significant difference in intertemporal choice behavior between unipolar and bipolar disorder patients. Therefore, we combined both unipolar and bipolar patients (referred to as "depressive patients", hereafter). Moreover, most patients with depression were under medical treatment with 7 several types of antidepressants/mood stabilizers. We did not, however, find any significant effect of the types of antidepressants/mood stabilizers on intertemporal choice. Therefore, we did not divide the patients into sub-groups according to the types of antidepressants they intake. The characteristics of these subjects are described in more detail in Table 1. Subjects with past or current illegal drug use, assessed with Mini-International Neuropsychiatric Interview (MINI, see below), were excluded from the present study, in order to avoid the influences of substance abuse on intertemporal choice. 2.2 Intertemporal choice task As a standard instrument measuring degrees to which participants discount delayed reward and loss, we conducted a face-to-face task developed by neuropsychopharmacologists Bickel and colleagues [20]. It is to be noted that we have also utilized this task in our previous studies [8,9,21]. First, participants were seated individually in a quiet room, and faced the experimenter across a table. After that, participants received the simple instruction that monetary rewards (or losses) in this experiment were hypothetical, but the experimenter wanted them to think as though they were real money. Then the participants were asked to choose between the card describing money delivered immediately (or paid immediately, in the loss condition) and the card describing money delivered after certain delay (or paid after a certain delay, in the loss condition). The left card viewed by participants indicated the amounts of money that could be received immediately (or that had to be paid immediately, in the loss condition), and the right card indicated 100,000 yen that could be received after a certain delay (or that had to be paid after a certain delay, in the loss condition. For the delay discounting tasks, monetary rewards (or losses) and the delay time were printed on 3×5 index cards. The 27 monetary amounts were 100,000 yen (about $1,000), 99,000 yen, 96,000 yen, 92,000 yen, 85,000 yen, 80,000 yen, 75,000 yen, 70,000 yen, 65,000 yen, 60,000 yen, 55,000 yen, 50,000 yen, 45,000 yen, 40,000 yen, 35,000 yen, 30,000 yen, 25,000 yen, 20,000 yen, 15,000 yen, 10,000 yen, 8,000 yen, 6,000 yen, 4,000 yen, 2,000 yen, 1,000 yen, 500 yen and 100 yen. The seven time delays were 1 week, 2 weeks, 1 month, 6 months, 1 year, 5 years and 25 years. The experimenter turned the 27 cards sequentially. The card started with 100 yen, up to 100,000 yen in the ascending order condition, or started with 100,000 yen, down to 100 yen, in the descending order condition. For each card, participants chose either the immediate or the delayed reward (or loss). The experimenter wrote down the first 8 delayed reward chosen in the descending gain condition, first immediate reward chosen in the ascending gain condition, first immediate loss in the descending loss condition, and first delayed loss in the ascending loss condition. The average of these results (gain condition and loss condition were calculated separately), were used as the points or subjective equality (hereafter called the indifference point, a subjective value of delayed gain/loss) in the following analyses. This procedure was repeated for each of the seven delays (for more detail, see Bickel et al., 1999). The four conditions (ascending gain condition, descending gain condition, ascending loss condition, and descending loss condition) were conducted randomly for each participant. These conditions of the order of the delay discounting tasks did not significantly influence the results. Because our aim was to compare the group difference between healthy controls and depressive patients, we presented estimated parameters of the q-exponential discounting (see the next section) for group median data of the indifference point at each delay. It should also be noted that when the analysis was performed at the individual level, essentially the same conclusions were obtained. After the determination of indifference points in the delay discounting tasks, we estimated parameters for the q-exponential discount function (i.e., kq and q in Equation 1, corresponding to impulsivity and inconsistency, respectively) for gain and loss, separately. For estimating the parameters, we conducted nonlinear curve fitting with the Gauss-Newton algorithm implemented in R statistical language (nonlinear modeling package). 2.3 Questionnaires Mini-International Neuropsychiatric Interview (MINI) To examine the participants’ drug abuse and alcohol abuse, we used MINI [22]. This scale is also a structured diagnostic scale, administered by well-trained psychiatric doctors when interviewing participants. Although this scale can divided into several parts for assessing specific psychiatric disorders, we used only 2 parts – those for assessing drug abuse and alcohol abuse. Beck Depression Inventory-II (BDI-II) To assess both healthy and depressive participant’s degrees of depressive tendency, we assessed Beck’s Depression Inventory (more specifically, BDI-II) [23,24]. This scale is a commonly used self-report scale that measures severity of depression ‘over the past week’. The scale consists of 21 items that describe core depressive symptoms, with 9 score being rated on a 4-point scale. In this study, we used the Japanese version of the BDI-II [25,26]. Demographic questionnaire In addition to the above measures, participants completed questionnaires about sex, age, history of smoking, and suicide attempt histories. These demographic data did not significantly affect intertemporal choice behavior in the present study. Therefore, we simply compared parameters in the q-exponential discount function for gain and loss between healthy controls and depressed patients. 2.4 Data analysis For statistically testing differences in estimated parameters in the q-exponential discount function at the individual level, we utilized t-tests. It is also to be noted that we examined the fitness of the discount models with AICc (Akaike Information Criterion with small sample correction) and observed that the q-exponential discount function best fitted among other models (i.e., exponential and hyperbolic functions) in both healthy controls and depressive patients. All statistical procedures were conducted with R statistical language (http://www.r-project.org/). Significant level was set at 5% throughout. 3. Results 3.1 Impulsivity and inconsistency in controls and depressive patients Demographic characteristics of depressive patients and healthy controls are presented in Table 1. We observed that BDI-II scores were higher for depressive patients than controls (p<0.05), verifying that our present populations were appropriate for the objectives of the present study. The estimated parameters in the q-exponential discount function for gain and loss at the group level are summarized in Table 2 (Table 2-1 for gain, Table 2-2 for loss). Fitted q-exponential curves for group median indifference points were presented in Fig. 1 (Fig. 1A for gain, Fig.2B for loss). We can see that depressive patients less dramatically discounted delayed outcomes in the distant future, compared to healthy controls, implying that depressive patients make more patient (less impulsive) intertemporal choice plans in the distant future. However, the patient plan does not imply that their actual intertemporal choice action is also patient (less impulsive), because it is possible for preference reversal to occur due to inconsistency in 10 intertemporal choice (see section 1.2). In order to investigate differences in both impulsivity at delay D=0, (i.e., impulsivity in intertemporal choice action) and inconsistency in delay discounting of gain and loss by controls and depressive patients at the individual level, we conducted t-tests on estimated kq (impulsivity at delay D=0) and q (time-consistency) between controls and depressive patients. Consequently, we found that depressive patients had significantly larger kq (impulsivity at delay D=0) and smaller q (time-consistency) for both gain and loss, in comparison to controls (ps<0.05). In other words, depressive patients were more impulsive and inconsistent in delay discounting actions for both gain and loss than healthy control subjects (see Table 2, for group data). Moreover, as noted earlier, there was no significant effect of demographic variables other than depressive status (e.g. sex, age, histories of suicide attempts, and the type of antidepressants) on the estimated parameters in the q-exponential discount function. 3.2 Gain-loss asymmetries Next, we plotted the estimated q-exponential discount rate defined in Equation 4 (for gain and loss) of healthy controls and depressive patients (Fig. 2). As can be seen from Fig.2, depressive patients were more impulsive (i.e., a larger discount rate) in the near future, but less impulsive (i.e., a smaller discount rate) in the distant future, in comparison to healthy controls. This indicates that depressive patients may experience more exaggerated "preference reversal" in their intertemporal choice between plans and actions. It is important to note that depressive patients' discount rate for future loss at the delay of several decades is almost zero, indicating that depressive patients had hypersensitivity to potential bad outcomes in the extremely distant future (a red curve, Fig 2B). In order to examine differences in impulsivity and inconsistency in intertemporal choice for gain and loss at the individual level, we conducted t-tests on estimated kq (impulsivity at delay D=0) and q (time-consistency) between gain and loss. We then observed that kq was significantly smaller for gain; while q was larger for gain (p<0.05), indicating that intertemporal choice for loss is less impulsive (in line with previous reports on the sign effect on a discount rate [12]) but more inconsistent (see Table 2 for group data). 4. Discussion 4.1 Impulsivity and inconsistency in depressive patients' intertemporal choice As far as we know, this study is the first to utilize the q-exponential discount model 11 inspired by Tsallis' non-extensive thermostatistics, in order to examine impairments in depressive patients' decision-making. We observed that depressive patients were more impulsive in intertemporal choice actions and more inconsistent in temporal discounting, in comparison to healthy subjects. Our results imply that (i) depressive patients may experience preference reversal more frequently and dramatically, than healthy subjects, (ii) depressive patients may be more sensitive to potentially harmful events occurring in the extremely distant future, than healthy subjects, and (iii) although depressive patients' plans about the distant future may be forward-looking, their intertemporal choice actions (occurring at delay D=0) may be more myopic than healthy controls. Future studies should examine whether subjects who tend to make infeasible future plans are more susceptible to depression. Another possibility is that neuronal changes associated with depression (e.g., hypoactivation of serotonergic systems) may induce time-inconsistency in temporal discounting. Concerning this possibility, several studies have indicated that time-inconsistency in temporal discounting may result from nonlinear distortion of psychological time [27-29]. With respect to biophysical mechanisms underlying serotonergic modulation of time-perception in intertemporal choice, it should be noted that (i) biophysical simulation studies indicate that nonlinear psychophysical effects on sensation is mediated by electrical coupling between neurons [30] and (ii) serotonin modulates neuronal coupling [31]. Therefore, future neuroeconomic and biophysical studies should examine whether a decrease in serotoninergic activities induces both distorted time-perception and time-inconsistent discounting behavior. As noted in the introduction, findings on the relationship between serotonin and discounting have been mixed. Mobini and colleagues' study [18] demonstrated a reduction in serotonergic activities resulted in a significant increase in a discount rate; while Crean and colleagues' study [19] reported no significant effect. It should be noted that in Mobini and colleagues' study, the time-range of the intertemporal choice task was short; while Crean and colleagues' study employed longer delays , i.e., about 1 year (note that both Mobin et al.'s and Crean et al's studies did not employ the q-exponential discount function but a simple hyperbolic function [10-12], and therefore the estimated simple hyperbolic discount rates in Crean et al's study were under the influences of both relatively longer and shorter delays). Our present results of the relationship between depression (associated with a reduction in serotonergic activities) and temporal discounting may resolve the discrepancy: a decrease in serotonergic activities may increase a discount rate at short delays; while decrease a discount rate at longer delays (see Figure 2). This analysis is impossible without utilizing the q-exponential discount 12 function. A recent study reported that intertemporal choice for other is more inconsistent than for self [9]. It is important to examine whether intertemporal choice for other by healthy subjects involves similar neuropsychological processes to intertemporal choice for self by depressive patients. Neurochemically, a previous study implied that reduced (nor)adrenergic activities (assessed with salivary alpha-amylase levels) are associated with impulsivity in intertemporal choice [32]. It should therefore be examined whether (nor)adrenergic activities are likewise related to consistency in temporal discounting. 4.2 Sign effects on intertemporal choice We observed that the signs of the outcomes (i.e., gain or loss) markedly affect both impulsivity and inconsistency in intertemporal choice (i.e., kq and q). Although several neuroeconomic studies examined the neural correlates of discounting delayed monetary gains [33], little is known regarding the neural processing underlying discounting of monetary loss. Future studies should examine which neuro-biophysical processes mediate the observed sign effects. 4.3 Limitations and future directions As noted, the depressive patients in the present study were medicated with several types of antidepressants/mood stabilizers. Therefore, it is not completely evident that the treatments did not dramatically affect the intertemporal choice behavior. Nevertheless, our present results may not totally be attributable to the effects of antidepressants, because (i) we did not observe the effects of the types of antidepressants on the parameters in the q-exponential discount function, and (ii) BDI-II scores were significantly higher in the depressive patients than the controls, indicating that the patients were still depressed enough to elicit depression-induced alteration in intertemporal choice (we carefully scheduled our experiment so that the patients may participate during their depressive phase). In order to further resolve these issues, future studies should examine the effects of antidepressants on intertemporal choice by healthy human subjects (or animals). References 1. Must A, Szabó Z, Bódi N, Szász A, Janka Z, Kéri S. (2006) Sensitivity to reward and punishment and the prefrontal cortex in major depression. J Affect Disord. 90(2-3):209-215. 2. Radford MH, Mann L, Kalucy RS. (1986) Psychiatric disturbance and 13 decision-making. Aust N Z J Psychiatry. 20(2):210-217. 3. Must A, Juhász A, Rimanóczy A, Szabó Z, Kéri S, Janka Z. (2007) Major depressive disorder, serotonin transporter, and personality traits: why patients use suboptimal decision-making strategies? J Affect Disord. 103(1-3):273-276 4. Lejoyeux, M., Arbaretaz, M., McLoughlin, M., & Ades, J. (2002). Impulse control disorders and depression. The Journal of Nervous and Mental Disease, 190, 310-314. 5. Kim, C., Lesage, A., Seguin, M., Lipp, O., Vanier, C., & Turecki, G. (2003). Patterns of comorbidity in male suicide completers. Psychological Medicine, 33, 1299-1309. 6. Schweighofer N, Tanaka SC, Doya K. (2007) Serotonin and the evaluation of future rewards: theory, experiments, and possible neural mechanisms. Ann N Y Acad Sci. 1104:289-300. 7. Cajueiro D.O. A note on the relevance of the q-exponential function in the context of intertemporal choices. Physica A 364 (2006) 385–388 8. Takahashi T, Oono H, Radford MH. (2007) Empirical estimation of consistency parameter in intertemporal choice based on Tsallis' statistics. Physica A 381: 338-342 9. Takahashi T (2007) A comparison of intertemporal choices for oneself versus someone else based on Tsallis' statistics. Physica A 385: 637-644. 10. Ainslie, G (2005) Précis of Breakdown of Will. Behav Brain Sci. 28(5):635-650; discussion 650-73. Review. 11. Mazur, J. E. (1987). An adjusting procedure for studying delayed reinforcement. In: Commons, M. L., Mazur, J. E., Nevin, J. A., & Rachlin, H. (Eds), Quantitative analysis of behavior: the effect of delay and of intervening events on reinforcement value. Vol. 5, Hillsdale, NJ: Erlbaum, 55-73. 12. Frederick, S., Loewenstein, G., & O' Donoghue, T. (2002) Time discounting and time preference: a critical review. Journal of Economic Literature. 40, 350-401. 13. Madden, G. J., Petry, N. M., Badger, G. J., & Bickel, W. K. (1997). Impulsive and self-control choices in opioid-dependent patients and non-drug-using control participants: drug and monetary rewards. Exoerimental and Clinical Psychopharmacology, 5, 256-262. 14. Madden, G. J., Bickel, W. K., & Jacobs, E. A. (1999). Discounting of delayed rewards in opioid-dependent outpatients: exponential or hyperbolic discounting functions. Experimental and Clinical Psychopharmacology, 7, 284-293. 15. Petry, N. M., Bickel, W. K., & Arnett, M. (1998). Shortened time horizons and 14 insensitivity to future consequences in heroin addicts. Addiction, 93, 729-738. 16. Kirby, K. N., Petry, N. M., & Bickel, W. K. (1999). Heroin addicts have higher discount rates for delayed rewards than non-drug-using controls. Journal of Experimental Psychology: General, 128, 78-87. 17. Samuelson, P. A. (1937). A note on measurement of utility. Review of Economic Studies, 4, 155-161. 18. Mobini S, Chiang TJ, Al-Ruwaitea AS, Ho MY, Bradshaw CM, Szabadi E. (2000) Effect of central 5-hydroxytryptamine depletion on inter-temporal choice: a quantitative analysis. Psychopharmacology (Berl). 149(3):313-318. 19. Crean J, Richards JB, de Wit H. (2002) Effect of tryptophan depletion on impulsive behavior in men with or without a family history of alcoholism. Behavioral Brain Research. 136(2):349-357. 20. Bickel WK, Odum AL, Madden GJ. (1999) Impulsivity and cigarette smoking: delay discounting in current, never, and ex-smokers. Psychopharmacology (Berl). 146: 447-454. 21. Ohmura Y, Takahashi T, Kitamura N, Wehr P. (2006) Three-month stability of delay and probability discounting measures. Experimental & Clinical Psychopharmacology. 14(3):318-328. 22. Sheehan DV, Lecrubier Y, Sheehan KH, Amorim P, Janavs J, Weiller E, Hergueta T, Baker R, Dunbar GC.(1998) The Mini-International Neuropsychiatric Interview (M.I.N.I.): the development and validation of a structured diagnostic psychiatric interview for DSM-IV and ICD-10. Journal of Clinical Psychiatry. 59 Suppl. 20:22-33 23. Beck AT, Erbaugh J, Ward CH, Mock J, Mendelsohn M (1961) An inventory for measureing depression. Archives of General Psychiatry 4 : 561 24. Beck AT, Young J (1979) Handbook of studies on depression. Journal of Nervous and Mental Disease. 167: 719 25. Hiroe T, Kojima M, Yamamoto I, Nojima S, Kinoshita Y, Hashimoto N, Watanabe N, Maeda T, Furukawa TA. (2005) Gradations of clinical severity and sensitivity to change assessed with the Beck Depression Inventory-II in Japanese patients with depression. Psychiatry Research. 135(3):229-235. 26. Matsumoto T, Yamaguchi A, Chiba Y, Asami T, Iseki E, Hirayasu Y. (2005) Self-burning versus self-cutting: patterns and implications of self-mutilation; a preliminary study of differences between self-cutting and self-burning in a Japanese juvenile detention center. Psychiatry Clin Neurosci. 59(1):62-69. 27. Takahashi, T. (2005). Loss of self-control in intertemporal choice may be 15 attributable to logarithmic time-perception. Medical Hypotheses, 65, 691-693. 28. Takahashi, T. (2006). Time-estimation error following Weber-Fechner law may explain subadditive time-discounting. Medical Hypotheses, 67, 1372-1374. 29. Takahashi, T. (2007). Hyperbolic discounting may be reduced to electrical coupling in dopaminergic neural circuits. Medical Hypotheses, 69(1):195-198. 30. Copelli, M., Roque, A.C. Oliveira R.F. and Kinouchi, O. (2002) Physics of psychophysics: Stevens and Weber–Fechner laws are transfer functions of excitable media, Physical Review E, 65, 060901 31. Burrell BD, Sahley CL, Muller KJ. (2002) Differential effects of serotonin enhance activity of an electrically coupled neural network. Journal of Neurophysiology 87(6):2889-2895. 32. Takahashi T, Ikeda K, Fukushima H, Hasegawa T. (2007) Salivary alpha-amylase levels and hyperbolic discounting in male humans. NeuroEndocrinology Letters. 28(1):17-20. 33. Kable JW, Glimcher PW. (2007) The neural correlates of subjective value during intertemporal choice. Nature Neuroscience. 10(12):1625-133. 16 Figure legends Fig. 1 Indifference points and fitted q-exponential discount function for gain (A) and loss (B). Black and red dots are indifference points at delays for healthy controls and depressive patients, respectively. Black and red curves are best-fit q-exponential discount functions for healthy controls and depressive patients, respectively. Note that in (B), the vertical axis is the unsigned (absolute) subjective value of delayed monetary loss. Fig. 2 Estimated q-exponential discount rate: kq/(1+kq(1-q)D) for gain (A) and loss (B). Black and red curves are q-exponential discount rates for healthy controls and depressive patients, respectively. 17 Table 1. Means and standard deviations for demographic variables for depressed patients and healthy people Sex (% men) Age (years) BDI-II Depressive patients Mean 55.18 43.83 20.52* SD 2.25 2.42 Healthy controls Mean 40 47.6 8.93 SD 3.53 1.12 *:Significantly larger than healthy controls (p<0.05). BDI-II: Beck Depression Inventory (high BDI-II scores indicate severe depression). 18 Table 2-1. Group data of estimated parameters in q-exponential discounting for gain kq (impulsivity) q (consistency) Depressive patients 0.0006099 -2.4252055 > < Healthy controls 0.0004064 -0.2132638 Note that larger kq and q indicate more impulsive intertemporal choice at delay D=0 and more consistent intertemporal choice, respectively. Table 2-2. Group data of estimated parameters in q-exponential discounting for loss kq (impulsivity) q (consistency) Depressive patients 0.0005576 -72.5 > < Healthy controls 0.0002696 -1.7333967 Note that larger kq and q indicate more impulsive intertemporal choice at delay D=0 and more consistent intertemporal choice, respectively. Fig.1A 19 20 Fig.1B 21 Fig2.A 22 Fig.2B
1912.02331
1
1912
2019-12-05T01:17:30
Investigation of ephaptic interactions in peripheral nerve of sheep using 6 kHz subthreshold currents
[ "q-bio.NC" ]
The objective of this work was to determine whether application of subthreshold currents to the peripheral nerve increases the excitability of the underlying nerve fibres, and how this increased excitability would alter neural activity as it propagates through the subthreshold currents. Experiments were performed on two Romney cross-breed sheep in vivo, by applying subthreshold currents either at the stimulus site or between the stimulus and recording sites. Neural recordings were obtained from nerve cuff implanted on the peroneal or sciatic nerve branches, while stimulus was applied to either the peroneal nerve or pins placed through the lower hindshank. Results showed that subthreshold currents applied to the same site as stimulus increased excitation of underlying nerve fibres (p < 0.0001). With stimulus and subthreshold currents applied to different sites on the peroneal nerve, the primary CAP in the sciatic displayed a temporal shift of -2.5 to -3 us which agreed with statistically significant changes in the CAP waveform (p<0.02). These findings contribute to the understanding of mechanisms in myelinated fibres of subthreshold current neuromodulation therapies.
q-bio.NC
q-bio
1 Investigation of ephaptic interactions in peripheral nerve of sheep using 6 kHz subthreshold currents James Hope1,2, Narrendar Ravi Chandra1, Frederique Vanholsbeeck2,3, Andrew McDaid1 1The Department of Mechanical Engineering, The University of Auckland, Auckland 1010, New Zealand 2The Dodd Walls Centre for Photonic and Quantum Technologies, Auckland 1010, New Zealand 3The Department of Physics, The University of Auckland, Auckland 1010, New Zealand Abstract The objective of this work was to determine whether application of subthreshold currents to the peripheral nerve increases the excitability of the underlying nerve fibres, and how this increased excitability would alter neural activity as it propagates through the subthreshold currents. Experiments were performed on two Romney cross- breed sheep in vivo, by applying subthreshold currents either at the stimulus site or between the stimulus and recording sites. Neural recordings were obtained from nerve cuff implanted on the peroneal or sciatic nerve branches, while stimulus was applied to either the peroneal nerve or pins placed through the lower hindshank. Results showed that subthreshold currents applied to the same site as stimulus increased excitation of underlying nerve fibres (p < 0.0001). With stimulus and subthreshold currents applied to different sites on the peroneal nerve, the primary CAP in the sciatic displayed a temporal shift of -2.5 to -3 µs which agreed with statistically significant changes in the CAP waveform (p<0.02). These findings contribute to the understanding of mechanisms in myelinated fibres of subthreshold current neuromodulation therapies. Keywords: Neuromodulation, ephaptic interactions, nerve cuff, peripheral nerve 1 Introduction Neuromodulation therapies provide an alternative treatment modality for several drug-resistant neurological conditions, including Parkinson's, epilepsy, and depression [1-6]. While neuromodulation therapies generally administer suprathreshold currents to modulate activity in target neural cells, a smaller number administer subthreshold currents to either suppress or promote neural activity through partial hyperpolarization or depolarization, respectively, of cell membranes. For example, chronic subthreshold cortical stimulation for epilepsy [7, 8]; a high frequency (10 kHz) variant of spinal cord stimulation for pain management [2]; the conditioning current which precedes the suprathreshold current in transcranial magnetic stimulation for motor cortex studies [9]; and, transcranial current stimulation for depression [6]. 1 2 Researchers have demonstrated this partial polarization effect on the vestibular system by administering subthreshold current, with a band pass filtered random noise waveform, to the mastoid processes and observing subject sway responses which were highly coherent with the applied current polarity and magnitude [10-12]. This coherence was explained by stochastic resonance, wherein subthreshold components of periodic stimulus and random noise stimulus sum to become suprathreshold [13, 14]. In two further studies [15, 16], when researchers applied a subthreshold current transcutaneously to the tibial nerve through electrodes placed proximal to the ankle, again with a band pass filtered random noise waveform, participants reported an increase in sensitivity to vibration applied to the foot. The authors of these latter two studies did not identify the underlying mechanism causing increased sensitivity, but postulated in one [16] that the applied current might increase synchrony of the sensory receptor action potentials in the tibial nerve due to augmentation of ephaptic interactions between active nerve fibres. In ephaptic interactions, spatiotemporal variations in the electric field around an active neural cell influences activity in nearby neural cells by altering their membrane potentials. These interactions can improve synchrony of activity within populations of neural cells or triggering of neural activity in subthreshold cells, and may be induced either by naturally occurring physiological effects [17-19] or by artificially increasing excitability using chemicals [20, 21] and subthreshold electrical current [18, 22, 23]. In myelinated nerve fibres, modelling studies predict ephaptic interactions alter the propagation velocities of action potentials in neighboring active fibres which improves synchrony [24-27], and an in vivo study on rat showed increased activity in response to an electrical stimulus when it was temporally coupled with a compound action potential (CAP) [28]. Increasing synchrony and localized triggering of new action potentials in peripheral nerves using a subthreshold current presents an exciting prospect because such a paradigm could improve the signal to noise ratio in peripheral nerve interfaces, aid physical rehabilitation after spinal cord injury and stroke, and provide insight into mechanisms of subthreshold current neuromodulation therapies. In the current study, we investigated augmentation of neural activity in hind limb of sheep in vivo by applying subthreshold, 6 kHz, sinusoidal currents to the peroneal nerve between stimulus and recording sites. This paradigm differs significantly from the in vivo rat study in [28] which evaluated changes in excitability induced by surrounding neural activity, and is more similar to the stochastic resonance experiments in [15, 16], though in the current study subthreshold currents were administered via nerve cuffs instead of transcutaneous electrodes, and neural recordings were acquired. Two evoked stimulus sites were employed, one via a nerve cuff implanted on the peroneal nerve, and the second via pin electrodes placed distal to the hock. We hypothesized that (1) administering subthreshold currents would increase excitability of the underlying fibres, and (2) this increased excitability would increase synchrony in the neural activity and alter the propagation velocities. 2 Materials and Methods 2.1 Experiment apparatus Nerve cuff electrode arrays were fabricated from stainless steel foil and silicon using the method described in [29]. The 28-channel nerve cuff contained a 2 rings of 14 electrodes, spaced 7 mm apart, with 0.46 x 3 mm active 2 3 area on each electrode, Fig. 1a. Electrodes were coated with poly(3,4-ethylenedioxythiophene):p-toluene sulfonate (PEDOT-pTS) to reduce the electrode-tissue contact impedance. The 2-channel nerve cuffs contained 2 electrodes, each 9.5 x 1 mm, spaced 6 mm apart, Fig. 1b. The 28 channel and 2 channel electrode arrays were glued (Smooth-On Sil-Poxy®) into elastomer cuffs with dimensions (L x O.D. x I.D) of 20 x 8 x 4.5 mm and 6 x 6 x 3 mm, respectively, and with a slit along one side to allow implant on the nerve . The 28-ch nerve cuff used for recording was connected to a headstage (INTAN C3314) via an adaptor (INTAN C3410), Fig. 1c. Neural recordings were acquired from each of the 14 electrodes on one electrode ring, with the electrodes of the other ring shorted together and used as the reference. Data were low pass filtered at 5 kHz for anti-aliasing, sampled at 20 kS/s, software notch filtered at 50 Hz to remove mains noise, then streamed via a USB interface board (INTAN C3100) to host PC and saved as .rhd files for processing later. Stimulation was administered either through the pins placed along the cannon bone, or through electrodes in the most distally implanted nerve cuff on the peroneal nerve, and was generated by a bench-top pulse stimulator (AM- systems 2100) when digitally triggered (National Instruments CompactRIO® NI9403), Fig. 1c. The digital trigger was recorded on the USB interface board (INTAN C3100). Subthreshold currents were generated using a PCB with parallel current-source circuits, developed by The EIT Research Group at University College London, and available for download from https://github.com/EIT-team. Currents were switched on and off using solid state relays (IXYS CPC1017N) controlled via digital lines (National Instruments CompactRIO® NI9403), Fig 1c. 2.2 Tissue preparation and handling All animal procedures were approved by The University of Auckland Animal Ethics Advisory Committee. In total, two Romney cross breed sheep, female, and weighing 61 and 67 kg, were used in in-vivo experiments. Anesthesia was induced using by intravenous injection, and maintained using a mixture of isofluorane, oxygen and medical air administered via endotracheal tube. At the conclusion of experiments subjects were euthanized. To implant nerve cuffs, once subjects were anesthetised, the left hind leg was extended and loosely fixed in position around the hock and fetlock joint. An incision was made down the posterior side of the thigh to expose underlying muscle, then the Semitendinosus and Biceps Femoris muscles were parsed apart, and the sciatic, tibial, and peroneal nerves were isolated by cutting away adipose tissue [30]. In subject 1, two 28-channel nerve cuffs were implanted adjacent to one another on the peroneal nerve, Fig. 1d. In subject 2, a 28-channel nerve cuff was implanted on the sciatic nerve, and four 2-channel nerve cuffs were implanted adjacent to one another on the peroneal nerve, Fig 1e. After implantation the muscle cavity was filled with physiological saline, preheated to 38 ºC, to cover the nerve cuffs and exposed nerve. A stainless steel pin was placed in this saline which connected via a 460 kΩ resistor to ground. In both subjects, two pins were placed 100 mm apart, subdermally, along the cannon bone to allow stimulation of distal sections of the peroneal nerve. 3 4 Electrodes through 14 1 Electrodes 15 through 28 Elastomer cuff (a) 1 15 14 28 Electrodes Elastomer cuff (b) (c) (d) (e) Figure 1: The 28 channel electrode array with two columns of 14 electrodes after PEDOT coating (left) and assembled into an elastomer cuff (right) (a), and four 2 channel electrode arrays before (left) and after (right) assembly into an elastomer cuff (b); black scale bars are 5 mm. A schematic of the experiment apparatus with solid lines for current controlled signals, dashed for digital lines, and dotted for analogue voltage recording (c), where STC = subthreshold current. Schematic of nerve cuff configuration in subject 1 (d) showing two 28-channel cuffs on the peroneal nerve, and in subject 2 (e) showing four 2 channel cuffs on the peroneal nerve and a 28-channel cuff on the sciatic nerve. 2.3 Experiment protocol The subthreshold currents were biphasic, 6 kHz, sinusoidal waveform, and, in all but one protocol, +/- 70 µA amplitude. A 6 kHz sinusoid was selected to be sufficiently above the frequency components of neural activity to be filtered out during signal processing, and because transient impedance studies have indicated that 6 kHz results in more resistive current across the node of Ranvier and less capacitive current across the myelin sheath than in the neighboring frequencies [31, 32]. The amplitude of the stimulus pulses were selected to be 50 % higher than that which produced onset of twitching of the lower hindshank and phalanges, respectively, for stimulus administered via nerve cuff on the peroneal nerve and pins apposing the cannon bone. Twitching onset was characterized by manually triggering stimuli and increasing amplitude while visually monitoring the hind limb. 2.3.1 Hypothesis 1: Subthreshold current contribution to excitation of fibres In subject 1, the contribution of subthreshold currents to excitation of fibres was tested by administering a pulse +/-0.3 mA, 50µs/phase, biphasic, square pulse stimulus every 500 ms to the distal 28-ch nerve cuff, Fig. 1d, through electrodes E11 and E23, Fig. 1a. In conjunction with this stimulus and on the same nerve cuff, no subthreshold current was applied for 5 seconds, then one subthreshold current was applied through E1 and E15 4 5 for 5 seconds, then a second subthreshold current -- in phase with the first - was added through E4 and E18 for 5 seconds, Fig. 1a. This protocol was performed for a total of 600 seconds. The in-phase nature of the two currents produces a sinusoid with a +/- 140 µA amplitude within the cuff. 2.3.2 Hypothesis 2: Pin stimulus In subject 1, augmentation of neural activity using subthreshold currents was investigated using stimulus applied through two pins inserted in the hindshank and apposing the cannon bone. Here, a +/-10 mA, 1 ms/phase, biphasic, square pulse was administered, while subthreshold currents were applied and recordings acquired from the peroneal nerve using the protocol described above in section 2.3.1. This protocol was performed for a total of 600 seconds with both stimulation pins inserted on the posterior side of the cannon bone, then repeated for a total of 600 seconds with the anodic pin inserted on the anterior side of the cannon bone. Finally, the protocol was repeated for a total of 100 seconds with the stimulus duration halved to 0.5 ms/phase to distinguish neural activity and stimulus artefacts in the neural recordings. 2.3.3 Hypothesis 2: Nerve cuff stimulus In subject 2, augmentation of ephaptic interactions in the presence of three subthreshold currents was investigated using stimulus applied through a nerve cuff. Here, a +/-0.3 mA, 100 µs/phase, biphasic, square pulse was administered every 500 ms to the distal most 2-channel nerve cuff on the peroneal nerve, while no subthreshold current was applied for 5 seconds, then subthreshold currents were added one at a time, for 5 seconds, to each of the three adjacent 2-channel nerve cuffs, Fig. 1e. This protocol was repeated for a total of 600 seconds. The adjacent nature of the 2 channel nerve cuffs means the sinusoid amplitude did not exceed +/- 70 µA at any point along the peroneal nerve. Finally, to verify the influence of subthreshold current amplitude on the CAPs, the above protocol was repeated for a total of 600 seconds with the subthreshold current amplitude reduced to +/- 30 µA. 2.4 Data processing Recorded data were processed in MATLAB (R2018b Mathworks). Data were parsed into 500 ms duration segments beginning with the digital trigger to the pulse stimulator, grouped into sets based on the number of applied subthreshold currents, and then notch filtered at 6 kHz to remove artefacts from the subthreshold currents. Signals with significant drift or corruption by artefacts, defined as having a mean value outside the bounds of - 100 to 100 µV in the temporal window 20 to 30 ms, were removed. Two metrics were used to evaluate changes in neural activity caused by the subthreshold currents: CAP amplitude, and CAP temporal shift. (i) CAP amplitude, defined as the difference between the maxima of the peak and the minima of the adjoining trough, was calculated for each of the data segments then compared for each subthreshold current condition using unpaired t-tests. (ii) CAP temporal shift was analysed by up-sampling the mean recorded waveforms for each current condition by a factor of 100 using cubic spline interpolation to increase temporal resolution, then comparing the upsampled waveforms using crosscorrelation. In addition, CAP amplitude and CAP temporal shift were evaluated by calculating the difference in mean recorded waveforms, and identifying points outside the +/-3 sigma noise threshold. 5 6 3 Results 3.1 Surgery and nerve cuff implantation The femoral artery crossed over the sciatic, tibial and peroneal nerve branches, Fig. 2a, which hampered isolation of nerve branches from surrounding adipose tissue, and implantation of the nerve cuffs. The peroneal nerve was larger in subject 1, at 4 -- 5 mm diameter, than in subject 2, at 3 mm diameter, which was one reason that compelled the use of different nerve cuff configurations between the two subjects Fig. 2b-c. Tibial nerve Peroneal nerve 28ch cuff Femoral artery Fetlock joint Stimulus pins Sciatic nerve Peroneal nerve Four 2ch cuffs Phalanges Lower hindshank Cannon bone Hock Figure 2: The sciatic, tibial, and peroneal nerve branches and femoral artery within the muscle cavity prior to nerve cuff implantation (a). One 28 channel nerve cuff implanted on the peroneal nerve in subject 1 (b), and four 2 channel nerve cuff implanted on the peroneal nerve in subject 2 (c). The hind limb fixed in place around the fetlock joint and hock, with stimulus pins visible in the lower hindshank (d). 3.2 Hypothesis 1: Subthreshold current contribution to excitation of fibres Twitching of the lower hindshank was observed in response to the stimulus pulse, Fig. 2d. With two subthreshold currents applied, sustained extension of the lower hindshank was observed, indicating prolonged application of a +/- 140 µA amplitude, 6 kHz sinusoid is sufficient to activate motor fibres in the peroneal nerve which innervate the gastrocnemius muscle. In all three current conditions -- none, one, and two subthreshold currents -- a CAP was observed between 0.3 and 0.8 ms, and with a peak at 0.4 to 0.45 ms, after commencement of the stimulus pulse. In electrode 8, which exhibited the largest CAP amplitudes, amplitudes of (mean +/- standard deviation): 1.58 +/- 0.33, 1.99 +/- 0.49, and 2.22 +/- 0.56 mV, respectively, were observed for none, one, and two subthreshold current conditions, Fig. 3a-c. Comparing these values with one another using unpaired t-tests (N = 300) produced two-tailed p values of < 0.0001, which is considered extremely statistically significant. Cross correlation produced lag values of 0 for both subthreshold current conditions. The difference in mean recorded waveforms were not analysed for these data. 6 7 Stimulus artefact CAP Stimulus artefact CAP Stimulus artefact CAP (a) (b) (c) Amplitude V m 1 0.5 ms Figure 3: CAPs recorded from the peroneal nerve in response to stimulus applied, 30 mm distally, to the same nerve. Individual voltage recordings (shaded lines) overlaid with the mean and standard deviation (bold lines with error bars) with none (a), one (b), and two (b) subthreshold currents applied show that the subthreshold currents contribute to excitation of nerve fibres when applied to the same region of nerve as the stimulus pulse. 3.3 Hypothesis 2: Pin stimulus Minor twitching of the phalanges were observed in response to the stimulus pulse. As was the case earlier, with two subthreshold currents applied sustained extension of the lower hindshank was observed. The 1 ms/phase, biphasic, square pulse produced a stimulus artefact with 2.5 ms artefact from the pulse, followed by 1.5 ms of ringing, Fig. 4a-c. The shorter duration, 0.5 ms/phase, pulse produced a stimulus artefact with 1.5 ms from the pulse followed by 1.5 ms of ringing, Fig. 4d. With both stimuli, CAPs were observed from 4.5 ms onwards, indicating the artefact and CAPs could be distinguished form one another. In both pin configurations and for all three subthreshold current conditions -- none, one, and two subthreshold currents -- neural activity was observed between 4.5 and 12 ms after commencement of the stimulus pulse, and contained multiple peaks and troughs with combined amplitudes in the range of 5 to 20 µV, Fig. 4b-c. CAP amplitudes were not analysed in the individual segments because they were obscured by noise. In the mean waveforms, where CAPs were visible, no changes in CAP amplitudes were identified above the 3σ noise threshold because of the large noise, of σ = 0.8 to 2.2 µV, relative to the CAP amplitudes, Fig. 4e-g. Cross correlation produced lag values of 0 for both subthreshold current conditions. These results indicate no changes in the CAP amplitude or temporal shift were identified outside the noise threshold. 7 8 Stimulus artefact 2 ms V µ 0 0 5 Stimulus artefact CAPs 2 ms V µ 0 2 (a) (b) Stimulus artefact CAPs Stimulus artefact CAPs Stimulus artefact CAPs Stimulus artefact CAPs (c) (d) Stimulus artefact CAPs σ = 0.8 µV (e) σ = 1.3 µV (f) σ = 1.0 µV σ = 1.2 µV σ = 1.8 µV (g) σ = 2.2 µV Figure 4: CAPs recorded from the peroneal nerve in response to stimulus applied approximately 300 mm distally through two pins apposing the cannon bone; with pins either on the posterior side of the cannon bone (a -- b), or the posterior and anterior side of the cannon bone (c -- d). The individual (shaded lines) and mean recordings (bold lines) show a large stimulus artefact between 0 and 4 ms (a). A close up of the mean recordings for none, one, and two subthreshold currents in red, green, and blue, respectively, show multiple CAPs with 5 to 20 µV amplitudes between 4.5 and 12 ms with a 1 ms/phase stimulus pulse (b-c), or 4.5 to 10.5 ms with a 0.5 ms/phase stimulus pulse (d). No changes in CAP amplitudes can be seen outside the 3σ noise threshold because of the large noise, of σ = 0.8 to 2.2 µV (e-g). 3.4 Hypothesis 2: Nerve cuff stimulus In all four current conditions -- none, one, two, and three subthreshold currents -- a stimulus artefact was visible between 0 and 0.35 ms, followed by a large CAP with a mean amplitude of 2.13 to 2.14 mV and a peak at 0.7 ms after commencement of the stimulus pulse, Fig. 5a. Two smaller, secondary CAPs followed, with mean amplitudes of 13 +/- 1 and 35 +/- 1 µV and peaks at 2.85 +/- 0.05 and 5 +/- 0.05 ms, respectively, Fig. 5b. Lastly, a long duration CAP between 10 and 200 ms was visible with an amplitude between 59 to 61.5 µV and a peak at 50.9 to 53.3 ms, Fig. 5c. Multiple spikes were visible within the 10 to 400 ms temporal range, each with amplitudes of 10 to 400 µV and durations of 0.1 to 1 ms. The number and temporal location of the spikes varied between individual data segments, and were suspected to be caused by movement artefacts and evoked action potentials within the muscle fibres. The primary CAP amplitude was (mean +/- standard deviation): 2.141 +/- 0.086, 2.136 +/- 0.088, 2.130 +/- 0.096, and 2.140 +/- 0.089 mV, respectively, for each of the four current conditions, Fig. 5a. Comparison of these values 8 9 using unpaired t-tests (N=300) produced two-tailed p values between 0.13 and 0.83, which are considered to be not statistically significant. For the one, two, and three subthreshold currents, lag values of -5, -5 and -6 were calculated from cross correlation of the upsampled waveforms within the temporal window of 0.5 to 2 ms, corresponding to a temporal shifts of -2.5, -2.5 and -3 µs. These temporal shifts were visible in the difference in mean recorded waveforms, where the one, two, and three subthreshold current conditions all showed a positive difference at 0.65 ms followed by a negative difference at 0.75 ms, each with amplitudes of 60 to 80 µV, Fig. 5d. Using the unpaired t-tests (N=300), both the positive and negative peak difference amplitudes produced statistically meaningful p-values of 0.02 or less, despite large uncertainty at these points of 280 to 350 µV. The amplitudes of the secondary CAPs and long duration CAPs were not analysed in the individual segments because they were obscured by noise. In the mean waveforms, where CAPs were visible, no changes in CAP amplitudes were identified above the 3σ noise threshold, Fig. 5e-f. Cross correlation of the mean waveforms produced lag values of 0 for all current conditions. Stimulus artefact Large primary CAP peak CAPs Small secondary CAP peaks (a) Long duration CAP peak Spikes (b) (c) V µ 0 0 0 1 V µ 0 0 2 V µ 0 0 2 0 0.5 1 1.5 2 4 6 Time (ms) 8 10 100 200 300 400 (d) (e) (f) V µ 0 5 V µ 0 1 V µ 0 1 0 0.5 1 1.5 2 4 6 Time (ms) 8 10 100 200 300 400 (g) V µ 0 5 0 0.5 1 1.5 2 Time (ms) Figure 5: CAPs recorded from the sciatic nerve in response to stimulus applied approximately 70 mm distally on the peroneal nerve with none (red), one (green), two (blue), and three (purple) subthreshold currents applied (a -- c). The individual (shaded lines) and mean recordings (bold lines) show a large stimulus artefact between 0 and 0.4 ms, followed a large primary CAP in the 0 to 2 ms window (a), two small secondary CAPs in the 2 to 10 ms window (b) and a long duration CAP coinciding with multiple, short duration spikes in the 10 to 400 ms window (c). Significant differences between the mean recordings are visible for the primary CAP indicating a temporal shift of the CAP in the presence of the subthreshold currents (d), whereas any differences in the secondary and long duration CAPs are below the noise threshold (e-f). The differences between the mean recordings in the primary CAP are significantly reduced with the subthreshold current is reduced from +/- 70 to +/- 30 µA. 9 10 When the amplitude of the subthreshold currents was reduced to +/- 30 µV, lag values of -1, 1 and 0 were calculated from cross correlation of the upsampled waveforms within the temporal window of 0.5 to 2 ms, corresponding to a temporal shifts of -0.5, +0.5 and 0 µs in the negative direction. These temporal shifts were again visible in the difference in mean recorded waveforms, although with markedly smaller amplitudes than previously, Fig. 5e. 4 Discussion The first hypothesis, that the subthreshold currents contribute to excitation of the underlying nerve fibres, was confirmed through two observations: (i) that two subthreshold currents applied to the same section of nerve produce sustained extension of the hindshank; and, (ii) that the CAP amplitude increases with application of one subthreshold current, and then increases further with two subthreshold currents, when applied to the same section of nerve. This result is expected given the summative effect of currents on neural membrane excitability, and agrees in principle with stochastic resonance experiments [13, 14], as well as the ephaptic interaction study on rat in vivo [28] where the subthreshold component was instead produced by a CAP. The second hypothesis, that increased excitability would increase synchrony in the neural activity and augment the propagation velocities, was confirmed with low confidence by two observations from the large primary CAP produced in subject 2: (i) a lag of -2.5 to -3 µs in the presence of subthreshold currents; and, (ii) a biphasic difference in the mean waveforms of around 140 µV peak-peak amplitude in the presence of subthreshold currents. While confidence in the lag values is limited by the use of upsampling with spline interpolation and because they represent an increase in propagation velocity of only 0.43 %, they do agree with the statistically significant biphasic difference in the mean waveforms characteristic of a negative temporal shift. It is not clear whether the observed change in CAP velocity agrees with modelling predictions in [24-27], that ephaptic interactions slow down the CAP, because the models did not consider external current. In the present study, it is conceivable that the partially excited membranes took less time to depolarize, which would produce the observed increase in CAP velocity. If this was the case, however, it is not clear why the lag values did not increase linearly with the number of subthreshold currents applied. Finite element modelling of nerve fibres may provide insight into the observed change in CAP velocity, as well as the best way to configure multiple subthreshold currents to augment such an effect. Limitations of this study are the low number of subjects (n=2), low temporal sampling rate (20 kHz), high noise, and small range of stimulus and subthreshold current amplitudes and frequencies investigated. 5 Conclusion In this study, we confirmed that administering subthreshold currents increases excitability of the underlying fibres, and, with low confidence, that this increased excitability changes the propagation velocity of neural activity in the underlying fibres. While more work needs to be done in this area to improve confidence in the results, these initial findings contribute to understanding of possible mechanisms of neuromodulation using subthreshold currents. 10 11 Disclosures The authors have no relevant financial interests in this article and no potential conflicts of interest to disclose. Acknowledgements The authors would like to thank David Holder, Kirill Aristovich and Enrico Ravagli, from the EIT Research Group at University College London, for fabrication of electrode arrays used in this study; Darren Svirskis from the School of Pharmacy, The University of Auckland for assistance with PEDOT coating electrodes; and staff at the Faculty of Medical and Health Sciences, The University of Auckland for their help with experiments. References Amon, A. and F. Alesch, Systems for deep brain stimulation: review of technical features. Journal of Neural Transmission, 2017. 124(9): p. 1083-1091. Verrills, P., C. Sinclair, and A. Barnard, A review of spinal cord stimulation systems for chronic pain. Journal of pain research, 2016. 9: p. 481-492. McClintock, S.M., et al., Consensus recommendations for the clinical application of repetitive transcranial magnetic stimulation (rTMS) in the treatment of depression. The Journal of clinical psychiatry, 2018. 79(1). Markert, M.S. and R.S. Fisher, Neuromodulation - Science and Practice in Epilepsy: Vagus Nerve Stimulation, Thalamic Deep Brain Stimulation, and Responsive NeuroStimulation. Expert Review of Neurotherapeutics, 2019. 19(1): p. 17-29. Cristancho, M.A., et al., Vagus Nerve Stimulation (VNS), in Psychiatric Neurotherapeutics: Contemporary Surgical and Device-Based Treatments, J.A. Camprodon, et al., Editors. 2016, Springer New York: New York, NY. p. 99-116. Philip, N.S., et al., Low-intensity transcranial current stimulation in psychiatry. American Journal of Psychiatry, 2017. 174(7): p. 628-639. Lundstrom, B., et al., Trial stimulation and chronic subthreshold cortical stimulation to treat focal epilepsy. Brain Stimulation: Basic, Translational, and Clinical Research in Neuromodulation, 2019. 12(2): p. 502. Lundstrom, B.N., et al., Chronic Subthreshold Cortical Stimulation to Treat Focal Epilepsy. JAMA Neurology, 2016. 73(11): p. 1370-1372. Valero-Cabré, A., et al., Transcranial magnetic stimulation in basic and clinical neuroscience: A comprehensive review of fundamental principles and novel insights. Neuroscience & Biobehavioral Reviews, 2017. 83: p. 381-404. Collins, J.J., et al., Noise-enhanced human sensorimotor function. IEEE Engineering in Medicine and Biology Magazine, 2003. 22(2): p. 76-83. Wuehr, M., J. Decker, and R. Schniepp, Noisy galvanic vestibular stimulation: an emerging treatment option for bilateral vestibulopathy. Journal of Neurology, 2017. 264(1): p. 81-86. Serrador, J.M., et al., Enhancing vestibular function in the elderly with imperceptible electrical stimulation. Scientific reports, 2018. 8(1): p. 336. Wiesenfeld, K. and F. Moss, Stochastic resonance and the benefits of noise: From ice ages to crayfish and SQUIDs. Nature, 1995. 373(6509): p. 33-36. Hänggi, P., Stochastic resonance in biology: How noise can enhance detection of weak signals and help improve biological information processing. ChemPhysChem, 2002. 3(3): p. 285-290. Breen, P.P., et al., Peripheral tactile sensory perception of older adults improved using subsensory electrical noise stimulation. Medical Engineering & Physics, 2016. 38(8): p. 822-825. 11 1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. 13. 14. 15. 12 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 29. 30. 31. 32. Breen, P.P., et al., A new paradigm of electrical stimulation to enhance sensory neural function. Medical Engineering & Physics, 2014. 36(8): p. 1088-1091. Jefferys, J.G., Nonsynaptic modulation of neuronal activity in the brain: electric currents and extracellular ions. Physiological Reviews, 1995. 75(4): p. 689-723. Shivacharan, R.S., et al., Self-propagating, non-synaptic epileptiform activity recruits neurons by endogenous electric fields. Experimental Neurology, 2019. 317: p. 119-128. Han, K.-S., et al., Ephaptic Coupling Promotes Synchronous Firing of Cerebellar Purkinje Cells. Neuron, 2018. 100(3): p. 564-578.e3. Arvanitaki, A., Effects evoked in axon by the activity of a contiguous one. Journal of Neurophysiology, 1942. 5(2): p. 89-108. Jasper, H. and A. Monnier, Transmission of excitation between excised non‐myelinated nerves. An artificial synapse. Journal of Cellular and Comparative Physiology, 1938. 11(2): p. 259-277. Fröhlich, F. and D.A. McCormick, Endogenous electric fields may guide neocortical network activity. Neuron, 2010. 67(1): p. 129-143. Faber, D., Field effects in the CNS play functional roles. Frontiers in Neural Circuits, 2010. 4(15). Capllonch-Juan, M., F. Kölbl, and F. Sepulveda. Unidirectional ephaptic stimulation between two myelinated axons. in 2017 39th Annual International Conference of the IEEE Engineering in Medicine and Biology Society (EMBC). 2017. Binczak, S., J. Eilbeck, and A.C. Scott, Ephaptic coupling of myelinated nerve fibres. Physica D: Nonlinear Phenomena, 2001. 148(1-2): p. 159-174. Reutskiy, S., E. Rossoni, and B. Tirozzi, Conduction in bundles of demyelinated nerve fibres: computer simulation. Biological cybernetics, 2003. 89(6): p. 439-448. Schmidt, H. and T.R. Knösche, Action potential propagation and synchronisation in myelinated axons. PLOS Computational Biology, 2019. 15(10): p. e1007004. Bolzoni, F. and E. Jankowska, Ephaptic interactions between myelinated nerve fibres of rodent peripheral nerves. European Journal of Neuroscience, 2019. 50(7): p. 3101-3107. Christopher, A.R.C., et al., Electrode fabrication and interface optimization for imaging of evoked peripheral nervous system activity with electrical impedance tomography (EIT). Journal of Neural Engineering, 2019. 16(1): p. 016001. Vasconcelos, B.G., et al., Origin and distribution of the ischiatic nerve in mixed-breed sheep. Brazilian Journal of Veterinary Research and Animal Science, 2014. 51(2): p. 102-110. Y Aristovich, K., Donega, M., Blochet, C., Avery, J., Hannan, S., Chew, D., S Holder, D., Imaging fast neural traffic at fascicular level with electrical impedance tomography: proof of principle in rat sciatic nerve. Journal of Neural Engineering, 2018. Hope, J., et al., Increasing signal amplitude in electrical impedance tomography of neural activity using a parallel resistor inductor capacitor (RLC) circuit. Journal of Neural Engineering, 2019. 12
1702.01568
1
1702
2017-02-06T11:13:15
Interpretation of Correlated Neural Variability from Models of Feed-Forward and Recurrent Circuits
[ "q-bio.NC" ]
The correlated variability in the responses of a neural population to the repeated presentation of a sensory stimulus is a universally observed phenomenon. Such correlations have been studied in much detail, both with respect to their mechanistic origin and to their influence on stimulus discrimination and on the performance of population codes. In particular, recurrent neural network models have been used to understand the origin (or lack) of correlations in neural activity. Here, we apply a model of recurrently connected stochastic neurons to interpret correlations found in a population of neurons recorded from mouse auditory cortex. We study the consequences of recurrent connections on the stimulus dependence of correlations, and we compare them to those from alternative sources of correlated variability, like correlated gain fluctuations and common input in feed-forward architectures. We find that a recurrent network model with random effective connections reproduces observed statistics, like the relation between noise and signal correlations in the data, in a natural way. In the model, we can analyze directly the relation between network parameters, correlations, and how well pairs of stimuli can be discriminated based on population activity. In this way, we can relate circuit parameters to information processing.
q-bio.NC
q-bio
Interpretation of Correlated Neural Variability from Models of Feed-Forward and Recurrent Circuits Volker Pernice 1,2 Rava Azeredo da Silveira 1,2,3,∗ 1 Department of Physics, ´Ecole Normale Sup´erieure, PSL Research University, 2 Laboratoire de Physique Statistique, Centre National de la Recherche Scientifique, Universit´e Pierre et Marie Curie, Universit´e Denis Diderot, 75005 75005 Paris, France 3 Princeton Neuroscience Institute, Princeton University, Princeton, NJ 08544, Paris, France USA ∗ corresponding author February 7, 2017 Abstract The correlated variability in the responses of a neural population to the repeated presenta- tion of a sensory stimulus is a universally observed phenomenon. Such correlations have been studied in much detail, both with respect to their mechanistic origin and to their influence on stimulus discrimination and on the performance of population codes. In particular, recurrent neural network models have been used to understand the origin (or lack) of correlations in neural activity. Here, we apply a model of recurrently connected stochastic neurons to in- terpret correlations found in a population of neurons recorded from mouse auditory cortex. We study the consequences of recurrent connections on the stimulus dependence of correla- tions, and we compare them to those from alternative sources of correlated variability, like correlated gain fluctuations and common input in feed-forward architectures. We find that a recurrent network model with random effective connections reproduces observed statistics, like the relation between noise and signal correlations in the data, in a natural way. In the model, we can analyze directly the relation between network parameters, correlations, and how well pairs of stimuli can be discriminated based on population activity. In this way, we can relate circuit parameters to information processing. 1 Author Summary The response of neurons to a stimulus is variable across trials. A natural solution for reliable coding in the face of noise is the averaging across a neural population. The nature of this averaging depends on the structure of noise correlations in the neural population. In turn, the correlation structure depends on the way noise and correlations are generated in neural 1 circuits. It is in general difficult, however, to tease apart the origin of correlations from the observed population activity alone. In this article, we explore different theoretical scenarios of the way in which correlations can be generated, and we relate these to the architecture of feed-forward and recurrent neural circuits. Analyzing population recordings of the activity in mouse auditory cortex in response to sound stimuli, we find that population statistics are consistent with those generated in a recurrent network model. Using this model, we can then quantify the effects of network properties on average population responses, noise correlations, and the representation of sensory information. 2 Introduction In the search for clues about the function of neural circuits, it has become customary to rely upon recordings of the responses of large populations of neurons. These measurements exhibit the concerted activity of neural populations in different conditions, such as presentations of different stimuli, summarized by stimulus-dependent, high-dimensional statistics. The lowest moment of these statistics are also the best characterized, namely the mean response of neurons, the variability of response of single neurons, and their pairwise correlations. With these statistics in hand, one can ask two questions: How are they generated in the neural population? What purpose, if any, do they serve? While the first question is mechanistic and the second is functional, the two are intimately linked. We focus here on these questions and on their connection. Along the mechanistic line of research, a number of network models have been proposed to account for the statistics in measurements as well as to determine the relationship between anatomical and physiological parameters, on the one hand, and observed dynamics, on the other hand. In this spirit, for example models of balanced networks have been proposed to explain asynchronous and irregular spike trains [1, 2, 3]. Mechanistic explanations for the origin of pairwise correlation include the influence of recurrent connections [4, 5, 6] and global fluctuations (e.g., from top-down afferents) [7, 8, 9, 10, 11, 12]. Reading the origin of correlations from the recorded activity in a population of neurons is, however, a difficult task [13, 14, 7, 15]. Along the functional line of research, the objective is to relate the statistics of neural responses to the function of neural circuits; for example, to elucidate the role of the statistics in the representation of sensory information. The relation between correlated variability in the population response and the accuracy of stimulus representation, in particular, has been the object of much study in recent years [16, 17, 18, 19, 20, 21, 22, 23, 24, 25]. One reason for the focus on correlations is their possible effect in suppressing noise along relevant dimensions [26, 2, 27, 28]. Recent work has also illuminated the importance of the origin of correlations for their structure as it relates to the effect on stimulus representation [29, 30, 31]. In the present paper, we follow a three-pronged approach to relate the possible mechanistic origins of correlation to recorded statistics in cortical populations [32], on the one hand, and to relate network mechanisms to the representation of information, on the other hand. First, we study models of neural populations using the framework of Poisson processes [33] and we identify signatures in the population statistics of different mechanistic motifs. Second, we examine data on cortical populations in the light of our model results, and we find that the measured statistics are consistent with those generated by a recurrent network of Poisson neurons. Third, we use an idealized model of a recurrent cortical population, in which we can manipulate mechanistic parameters, to evaluate how the latter affect the accuracy of the representation of information. We find that, if correlations are not too weak, correlations generated within recurrent networks can be distinguished from those that are due to external 2 signals. The estimated parameters from the data suggest that the interpretation of neural spike trains as Poisson processes implies a strong amplification together with noise generation by the network dynamics. 3 Materials and Methods 3.1 Experimental data set The data set was first published and analyzed in [32]. It consists of the activity of neural populations (46-99 neurons) recorded using calcium imaging in the auditory cortex of mice. Animals were isoflurane anesthetized (1 %). Signals were obtained from neurons labeled with the synthetic calcium indicator OGB1. Fluorescence was measured at 30 Hz sampling rate, and firing rates inferred from temporal deconvolution of the fluorescence signal. Up to 6 neural populations in each of 14 animals were recorded. The data points we use are the average firing rates over a window of 250 ms after presentation of each of 65 different sound stimuli, each for 15 trials. Responses were measured relative to spontaneous activity. Hence, negative responses occurred, if the stimulus-evoked firing rates were smaller than the spontaneous ones. The stimuli consisted of a range of different pure tones recorded for different sound intensities, as well as a number of natural sounds. We do not distinguish individual stimulus identities, but instead regard the set of stimuli as a diverse and, to some degree, generic ensemble. 3.2 Description of response variability If the vector of population activity in trial T of stimulus s is r(s, T ), the average response across trials for this stimulus is denoted by r(s) = (cid:104)r(s, T )(cid:105)T . To measure covariability across trials in a given stimulus condition between neurons i and j, we use the noise covariances, defined as Cij(s) = cov(ri(s, T ), rj(s, T ))T = (cid:104)ri(s, T )rj(s, T )(cid:105)T − (cid:104)ri(s, T )(cid:105)T(cid:104)rj(s, t)(cid:105)T . (1) A measure of the strength of pairwise noise correlations across stimuli is the average correla- tion coefficient, Here, (cid:104).(cid:105)s denotes the average over all stimuli presented. This quantity is to be contrasted with the signal correlation, (2) (cid:42) (cid:43) . s Cij(s) (cid:112)Cii(s)Cjj(s) cov(cid:0)ri(s), rj(s)(cid:1) (cid:113) var(cid:0)ri(s)(cid:1) svar(cid:0)rj(s)(cid:1) s cN ij = cS ij = , s (3) which measures the similarity of average responses across stimuli. As a measure of how the orientation of the high-dimensional distributions changes across stimuli, we use the variance of the population activity projected along the direction of the normalized mean response, ¯r(s) = r(s)/r(s), namely, σ2 µ(s) = Cij(s)¯rj(s)¯ri(s). (4) (cid:88) ij 3 √ This quantity can be compared to the variance projected along the diagonal direction, ¯d = (1, . . . , 1)T / N , (cid:88) ij σ2 d(s) = Cij(s) ¯di ¯dj, µ and σ2 d are normalized by the sum of the variances, σ2 which corresponds to a uniform averaging of the covariances. To compare these quantities across stimuli, σ2 i Cii(s). If, for a given stimulus, all neurons are equally active on average, then ¯r = ¯d and σ2 d = σ2 µ. As we show, different circuit models predict different stimulus dependencies of σ2 µ and σ2 d. These differences are most apparent for the stimuli for which the average population response differs strongly from a uniform population response. A deviation from a uniform response can be measured by the angle between mean response and diagonal, or cos(d, r) = ¯rT · ¯d. For a graphical illustration of these measures, see Fig. 1. (5) all(s) = (cid:80) Figure 1: Properties of response distributions and network scenarios. A: Examples for distributions where the variability is to a large degree in the diagonal direction (response distribu- tion indicated by blue ellipse), resulting in a large value of σd, and where the variability is mainly in the direction of the average response (red ellipse), resulting in a larger value of σµ. B: Response distributions for two stimulus ensembles. Either σµ (red set) or σd (blue set) remains large across stimuli. A dependence as in the red set appears in a model of shared gain fluctuations, while a dependence as in the blue set may arise in densely connected recurrent or feed-forward networks. C: Different network layouts that induce correlated activity. Connections (arrows) to and be- tween neurons (dots) vary in strength. Dashed arrows indicate multiplicative effects on firing rates. 4 3.3 Recurrent model of Poisson neurons with noisy input (cid:90) ∞ ri(t) = rext,i(t) + We model an intrinsically noisy process of spike generation in neurons, in which the effect of presynaptic spikes on neural activity is captured through linear modulations of an underlying firing rate. Specifically, the spike train of neuron i is a realization of an inhomogeneous Poisson process with a time-dependent firing rate, ri(t), calculated as gij(τ )sj(t − τ )dτ, (cid:88) where rext,i(t) is the external input, sj(t) =(cid:80) t(cid:48) δ(t−t(cid:48)) is the recurrent input from presynaptic spike trains, and the causal coupling kernel, gij(τ ), defines the interactions between neurons in the network. The external input is an analog quantity which can be viewed as resulting from a convolution between a linear filter and an incoming spike train from "non-local" presynaptic neurons. We treat it as a noisy, but stationary, signal, i.e., its mean and higher moments do not depend on time. While this model of interacting point processes [33] defines the full temporal dynamics of the system, we are only interested in time-averaged rates or, equivalently, spike counts. Their expected values and (co-)variances across realizations can be obtained if the coupling matrix is known. (6) 0 j Assuming stationarity of input spiking and firing rates, we can solve for the average rate vector, r = (r1, . . . rN ), across trial realizations, as Here, I denotes the identity matrix, G is the steady-state coupling matrix with elements and rext is the average input vector. To describe the correlation in the activity, we consider the spike count in a long time bin, ∆, defined as The total input in the time bin is given by ni(∆) = si(t)dt. (9) (7) (8) (11) (12) r = (I − G)−1rext. Gij = gij(τ )dτ, (cid:90) ∞ 0 (cid:90) t+∆ (cid:90) t+∆ t rext(∆) = (10) and is a random variable with a normalized variance, Vext ≡ var(rext(∆))/∆ (across choices of different time bins). From spectral properties of the covariance [33] and the relation between covariance and count correlations [34, 5], it can be shown that the count covariance normalized by the time bin duration, C, with elements rE(t)dt, t cov(cid:0)ni(∆), nj(∆)(cid:1) , Cij ≡ lim ∆→∞ ∆ C = (I − G)−1(cid:0)D[r] + D[Vext](cid:1)(I − GT )−1. is given by the matrix The term D[r] denotes a diagonal matrix with elements D[r]ij = riδij, and similarly D[Vext]ij = δijVext,i. The effect of the recurrent network interactions is reflected in the transfer matrix, B ≡ (I − G)−1. 5 (13) In Eq. (12), one component of the covariance arises from the variance of the external input, Vext. Although the intrinsic noise from the spike-generating Poisson process is dynamical, its effect is that of an additive contribution to the external variance (see also [35]). The dependence on the rate vector, in Eq. (12), results from the Poisson character of the noise, in which the variance equates the mean. In the experimental data set, firing rates are measured relative to spontaneous activity. To take this into account, an offset, a, can be added to the rates in Eq. (12). Henceforth, we examine the model through Eqs. (7) and (12) for the mean neural activities and their covariances, respectively. 3.3.1 Random effective network and stimulus ensemble The rates and covariances given in Eqs. (7) and (12) depend on the recurrent coupling matrix within the network, G, only via effective couplings defined by the transfer matrix, B = (I − G)−1. In numerical calculations, we choose the entries of G independently from a normal distribution with positive mean; this affords a specific structure to the matrix B. The empirical mean and variance of the elements of this matrix will be called (cid:104)B(cid:105) and var(B). In a number of analytical calculations, for the sake of simplicity, we neglect the specific structure of B and model it as a random matrix with elements independently drawn from a normal distribution with corresponding mean and variance. The elements of B are not direct, but ef- fective connections, because they reflect also the effects of indirect connections in the network, hence the expression "random effective network." The variability of effective connections can be quantified by the ratio ρ = var(B)/(cid:104)B(cid:105)2. The stimulus ensemble is modeled as a set of random vectors, rext(s). We assume that the elements of these vectors are independent across neurons and stimuli, and normal, and we call the mean and variance of the corresponding normal distribution (cid:104)rext(cid:105) and var(rext). The relative variability of the inputs is defined as ρE = var(rext)/(cid:104)rext(cid:105). This variability is high if both excitatory and inhibitory inputs are present, and, together with the network parameters, it determines the variability of the average responses across stimuli. In numerical calculations, we set Vext,i(s) = rext,i(s), as dictated by a Poisson process. We allow for input signal correlations, i.e., the average inputs to a pair of neurons can come with a non-vanishing correlation coefficient, cin = (cid:104)rext,i(s)i, rext,j(s)(cid:105)s. 3.4 Models of correlated activity from shared inputs or gain modulations Correlations in the activity of neurons can have other origins besides recurrent connections. In parallel with the recurrent network model, we consider two alternative prototypical models in which correlations originate from shared inputs or from shared gain fluctuations, respectively. The three different scenarios are illustrated in Fig. 1C. In Appendix 8.2, we show how, formally, these models can also be cast as special cases of the recurrent model. 3.4.1 Feed-forward model with shared input We consider a simple, two-layer, feed-forward network, in which N output neurons receive inputs from N independent presynaptic neurons with spike trains sj(t). The contributions of the input neurons to the firing rates of the output neurons are determined by feed-forward connection kernels fij(τ ), so that the observed firing rates are 6 (cid:90) ∞ (cid:88) ri(t) = fij(τ )sj(t − τ )dτ. (14) 0 j We define the feed-forward coupling matrix F by its elements, Fij = (cid:82) ∞ As we show in Appendix 8.2, this scenario represents a special case of the framework described in Section 3.3, so that firing rates and count covariances can be calculated correspondingly. fij(τ )dτ . With the time-averaged firing rates of the external input for stimulus s, rext(s), the average spike counts of the output neurons are given by 0 r(s) = F rext(s). (15) If the spike trains of the input neurons are Poisson processes, the input variance is equal to the rate, Vext = rext, but we will allow for more general inputs. The count covariances are then given by C = F D[Vext]F T + D[r]. (16) We denote again by D[r], D[Vext] diagonal matrices with diagonal elements given by the vectors r, Vext. The first term describes covariances resulting from the combined shared inputs. The second term results from the contribution to the count variances of the Poisson spike generation in output neurons, which is independent across neurons. If rates are observed only up to a stimulus-dependent offset, a, Eq. (16) has to be replaced by C = F D[Vext]F T + D[r + a]. (17) This scenario can be compared directly to the random effective recurrent network model, if one uses F = B for an identical ensemble of stimuli (see Section 3.3.1). Comparing Eq. (17) to Eq. (12), we note that the difference between the two models is that the internally generated noise in the feed-forward model, D[r + a], is uncorrelated and contributes only to the variances, while in the recurrent network it is filtered by the network and therefore correlated. 3.4.2 Model with shared gain fluctuations In addition to recurrence and shared inputs, shared gain fluctuations have also been proposed as a source of correlation in neural populations by various authors [7, 36, 24, 10]. In this model, the firing rate of the neural population is given by the product of a constant, stimulus dependent vector, r(s), and a scalar fluctuating signal, (t). We set the time-averaged value of the fluctuations to unity, (cid:104)(t)(cid:105)t = 1, so that the time-averaged firing rate is r. The resulting count covariance matrix is given by C = D[r] + rrT Vext, (18) where the scalar variance, Vext, reflects the strength of the variation of the fluctuating signal. In other words, the firing rate of a Poisson neuron is modulated multiplicatively by a fluctu- ating "gain" signal. One consequence is that the covariance of a neuron pair is proportional to the product of their firing rates. The general model described in Section 3.3 reduces to this model if for any given stimulus the activity of the population results from a single input neuron projecting to the output neurons with corresponding weights (see Appendix 8.2). Here, the average responses can be chosen freely. When comparing this scenario to the recurrent and feed-forward scenarios, we use the ensemble of average responses resulting from the corresponding network scenario. If the stimulus dependence of the firing rates is measured relative to an offset a, the covariances are given by C = D[r + a] + (r + a)(r + a)T Vext. (19) 7 3.5 Stimulus discriminability (cid:16) (cid:17)−1(cid:16) (cid:17) r(s1) − r(s2) In order to evaluate the influence of correlation on neural coding, we examine the discrim- inability of a set of stimuli from the population activity. In the models, a stimulus, s, evokes activity characterized by an average response vector, r(s), and a covariance matrix, C(s). We seek a simple measure for evaluating the possibility of attributing a given population re- sponse to one of two discrete stimuli, s1 and s2, unambiguously. For this, we assume that the high-dimensional distributions of responses are Gaussian, and project the two distributions on a single dimension where we calculate the signal-to-noise ratio as our measure. The best projection follows from Fisher's linear discriminant analysis: the most informa- tive linear combination of neuron responses, denoted by w · r, is achieved if the vector w points in the most discriminant direction, w = C(s1) + C(s2) . (20) The mean and variance of the projected distributions onto the normalized direction, ¯w, are ¯wT r(si) and σ2 as si = ¯wT C(si) ¯w, for i ∈ {1, 2}. We can then define the "signal-to-noise ratio," (cid:12)(cid:12) ¯wT ·(cid:0)r(s1) − r(s2)(cid:1)(cid:12)(cid:12) . (21) S = σs1 + σs2 Larger values of S correspond to better discriminability. To quantify the effect of correlation on discriminability, we compare the quantity S, defined in Eq. (21), with the quantity Sshuffled obtained from a "shuffled data set", in which responses are shuffled across trials (in experimental data) or off-diagonal covariance elements, Ci(cid:54)=j(s), are set to 0 (in models). A ratio Sshuffled/S smaller than unity indicates that correlation is beneficial to discriminability. We note that S is an approximation for a measure based on the optimal linear classifier, for which the threshold separating the two one-dimensional projected distributions has to be calculated numerically. A measure akin to S used in similar contexts is the linear Fisher information [16], valid for continuous stimuli. An advantageous property of S is its invari- ance under linear transformations: if all responses, r, are fed into another network whose output, Br, results from a product with an invertible matrix B, S does not change. This obtains because the mean responses are transformed into Br(s), while the covariances are transformed into BC(s)BT . Intuitively, a simple matrix multiplication is accommodated for by a corresponding change in the most discriminative direction w on which we project. 3.6 Two-population model In order to develop an intuitive understanding of the way in which correlations generated in a recurrent network influence coding, we examine a highly simplified model. The recurrent network consists of two excitatory sub-populations, labeled E and E(cid:48), each made up of N neurons. We ask to what extent their activity discriminates two stimuli when these elicit preferential responses in the two sub-populations, respectively. The activity in the network is determined by Eqs. (7) and (12). For the sake of simplicity, we can set the external variance to zero, such that the input to the network is fully defined by its mean input, rext. We assume some further simplifications, for the sake of calculational ease: each neuron in sub-population L projects to a fixed number, nKL, of neurons in sub- population K; all non-zero coupling weights are identical, Gij = gE. Additionally, all neurons in the same population receive identical external input. Then, the two-component vectors Rext(s1) = (1 + ∆, 1)T and Rext(s2) = (1, 1 + ∆)T define a pair of stimuli. Each component denotes the input to each neuron in the corresponding population. 8 We reduce the dimensionality of the system by considering the population activity on a macroscopic level. The average response of population K across trials is defined as RK = rk, (22) and its trial-to-trial variability is described by the population covariance matrix, with elements defined as ΣKL = Ckl. (23) It turns out that these macroscopic quantities depend only on the overall number of connec- tions between populations, not on a specific network realization. For example, the sum of the rates of neurons in one population depends only on the sum of the inputs to all neurons, but not on how these inputs are distributed among the neurons. This is a consequence of the linearity of the dynamics and the assumption that each neuron has a fixed number of output connections, as we show in detail in Appendix 8.4. This idea was already applied in Refs. [5, 6]. More precisely, we define the coupling within each population, Γs = nEEgE = nE(cid:48)E(cid:48)gE, and across populations, Γc = nEE(cid:48)gE = nE(cid:48)EgE, which make up a population coupling matrix (cid:88) (cid:88) k∈K k∈K,l∈L (cid:18)Γs Γc (cid:19) Γc Γs Γ = . (24) Applying the definition in Eqs. (22) and (23), we can rewrite the microscopic Eqs. (7) and (12) as population equations, R(s) = (I − Γ)−1N Rext(s) (25) and Σ(s) = (I − Γ)−1D[R(s)](I − Γ)−1. (26) The population transfer matrix is defined by P = (I−Γ)−1. Up to a factor N , these equations are equivalent to the microscopic equations in the restricted context of the two-population model. 4 Results We discuss the characteristics of the noise statistics that emerge from neural dynamics in models of recurrent networks, feed-forward networks, and networks with global gain fluctua- tions. We also look for coarse statistical signatures of each of these three model structures, which can then be compared with data. In this perspective, we analyze cortical recordings in mouse. These recordings were made in the auditory cortex, while the animal was stimulated with musical tones [32]. We then discuss the implications of the neural noise statistics for stimulus encoding. Finally, we present a highly simplified model of network dynamics, paired down to include only a few parameters, in order to develop an intuition of the link between network structure and stimulus encoding. 4.1 Population signatures of noise statistics in a recurrent net- work model One population feature which distinguishes the possible mechanisms of generation of noise correlation is the relation between the population-averaged response, (cid:104)r(cid:105)(s) = (cid:104)ri(s)(cid:105)i = 9 1 N (cid:80) Cij =(cid:80) i ri(s), on the one hand, and the population-averaged variance, (cid:104)Cii(cid:105)i(s) = 1 i Cii(s), or the noise covariance averaged across pairs (cid:104)Cij(s)(cid:105)i(cid:54)=j = i(cid:54)=j Cij, on the other hand. In this section, we analyze these relations for Poisson neurons in a recurrent network with random effective connections, as described in Section 3.3.1. Using Eq. (12), which reads N (N−1) N 1 k BikBjk(rk(s) + a + Vext,k) for the pairwise covariances, we derive the expression (cid:80) (cid:80) (27) (28) for the average variance, and the expression (cid:104)Cii(cid:105)i(s) ≈ N(cid:104)B2(cid:105)(cid:16)(cid:104)r(cid:105)(s) + a + (cid:104)Vext(cid:105)(cid:17) (cid:104)Cij(cid:105)i(cid:54)=j(s) ≈ N(cid:104)B(cid:105)2(cid:16)(cid:104)r(cid:105)(s) + a + (cid:104)Vext(cid:105)(cid:17) (cid:80) (cid:80) N 2 N 2 ij B2 for the average covariance. (See Appendix 8.3.4 for mathematical details and Fig. 2 for nu- merical results.) The quantity (cid:104)B(cid:105) = 1 ij Bij denotes the average strength of the effective connections in the network across neuron pairs and, correspondingly, (cid:104)B2(cid:105) = 1 ij the average of their square. The variance of the input, averaged across neurons, is (cid:104)Vext(cid:105), and a denotes a potential constant offset in the observed firing rates. The variance of a Poisson variable is proportional to its mean. This contribution from the Poisson spike generation is reflected by the term (cid:104)r(cid:105) in the two equations and is the reason for the linear relation between average response and average variance/covariance - for non-Poisson spike generation, these relations would be non-linear. As we see, not only the external noise, (cid:104)Vext(cid:105), but also the internally generated noise are amplified by the recurrent connections. The amplification factor, or the slope of the linear relation, N(cid:104)B(cid:105)2 in Eq. (28), is smaller for the covariances, where it depends only on the mean strength (cid:104)B(cid:105) of the effective connections. For the variances, expressed in Eq. (27), the averaging in the slope N(cid:104)B2(cid:105) is carried out on the square of the effective connections. The slope is therefore larger, and depends both on the average strength of connections and their variability. In the relations for both average population variance and covariance, the ratio between the intercept and the slope is (cid:104)Vext(cid:105) + a. Up to a possible offset, a, that can result from a measurement of firing rates relative to a constant baseline activity, it represents the strength of the external noise. In summary, in the recurrent network model, both internally generated noise and external noise are amplified (or attenuated) by the recurrent connections, and the parameters in the relationship between (co-)variances and population response can be related to network and input parameters. 10 Figure 2: Variability in the recurrent network model. A: Relation between response and variance averaged across population for randomly chosen stimulus vectors (each dot one stimulus). Four networks are generated with normally distributed connection strengths. Larger ρ, calculated from effective connections in the transfer matrix B, indicate larger variability of effective weights. Lines indicate analytic predictions, Eqs. (27) and (28). B: Same for covariances, averaged across all neuron pairs. C: The ratio of slope and intercept of linear fits corresponds to the strength of the input variance, which is identical for all networks, and is consistent for variances and covariances. See Appendix 8.1 for further details and the numerical parameters. 4.2 Population signatures of noise statistics in feed-forward network models In a feed-forward network, both shared input units and global gain fluctuations can yield noise correlations [7, 10, 24]. Here, we compare the population signatures of the noise statistics in the case of these two alternatives with those from a recurrent model. This serves to point out differences between the outputs of the various models as well as to motivate the data analysis to which we turn below. The three scenarios differ in the scaling of the population-averaged covariance, (cid:104)Cij(cid:105)i(cid:54)=j(s), with respect to the population response, (cid:104)r(cid:105)(s) (Fig. 3). We already demonstrated in the pre- vious section that both variances and covariances scale linearly with the population response in the recurrent network. In a feed-forward network, pairwise covariances are given by Eq. (16), which can be rewritten as (cid:88) Cij = δij(ri + a) + FikFjkVext,k. (29) k Note that, in contrast to the recurrent network, the neural firing rates, ri, only affect diagonal entries, and therefore the variances. The average covariance can be expressed as (cid:104)Cij(cid:105)i(cid:54)=j(s) ≈ N(cid:104)F(cid:105)2(cid:104)Vext(cid:105)(s), (30) and does not directly depend on the average population response, (cid:104)r(cid:105)(s) (see Appendix 8.3.3). If there are excitatory as well as inhibitory inputs, also the average input variance, (cid:104)Vext(cid:105), can be uncorrelated to the population response: for Poisson external input, in each input channel i the variance is as large as the mean input, Vext,i = rext,i. Nonetheless, because the variance is always positive, including for negative input, the average variance (cid:104)Vext(cid:105) can be 11 decorrelated from the average input, (cid:104)rext(cid:105), and thus from the average output, (cid:104)r(cid:105). Intuitively, because inhibitory channels add variance, but decrease the average input, the combined input does not have to be proportional to the combined variance. The extent to which this is true depends on the balance between positive and negative elements in rext, which we measure by the input variability ρE = var(rext)/(cid:104)rext(cid:105). In a model network with global gain fluctuations, from Eq. (18), the covariances are given by Again averaging only across the off-diagonal elements, one finds that the average covariances scale quadratically with the population response, as Cij = δijri + rirjVext. (31) (cid:104)Cij(cid:105)i(cid:54)=j = Vext(cid:104)r(cid:105)2. (32) This is a direct consequence of the fact that pairwise covariances are proportional to the product of the firing rates of the pair of neurons, which holds for large N independent of the distribution of average responses. By contrast, as is apparent from the expression for Cij, the individual variances and the population averaged variance each have both a quadratic and a linear contribution in r (and (cid:104)r(cid:105), respectively) (see also e.g. [7]). (Further details are provided in Appendix 8.3.) Until now, we examined how the average covariance (cid:104)Cij(cid:105)i(cid:54)=j changes across stimuli. Fur- ther differences between the different scenarios can be related to the detailed structure of C, which determines the shape of the response distributions. By shape, we refer to the ge- ometric orientation and extent of the multi-dimensional ellipsoid cloud that corresponds to the response distribution in the space spanned by the responses of the individual neurons. A two-dimensional sketch of such an ellipse, and the geometric interpretation of the quantities we use is depicted in Fig. 1A (see also Methods 3.2). In words, the way the orientation of the response distribution changes with the average response varies between the three scenar- ios: the long axis of the ellipsoid indicating the direction of the largest variance is aligned along the diagonal for the feed-forward and the recurrent network models, but in the direc- tion of the mean response for the gain fluctuation model (Fig. 1B). To capture this relation quantitatively, we consider the projected variances, σ2 d (see Fig. 3 and methods for details). There, the "diagonal direction" corresponds to a response where all neurons are equally active. µ and σ2 Quantitatively, we find that in the recurrent and feed-forward network models with rela- tively strong correlations, there is always a large fraction of variance in the diagonal direction, independent of the stimulus. The variance projected along the direction of the average re- sponse is only large if the latter happens to be similar to the diagonal direction. By contrast, in the gain fluctuation model, the variance projected along the direction of the average re- sponse is approximately constant across stimuli. This behavior of the gain fluctuation model is due to the fact that the variability across trials is multiplicative on the mean response, and hence variability is large in this direction for all stimuli. In the network models, strong correlations result from effective interactions between neu- rons or from shared input. Each neuron in the recurrent network, or each input channel in the feed-forward network, is effectively connected to a large part of the remaining neurons and contributes to their correlations. Because our networks are not balanced, these contributions are mostly positive, even for inhibitory input channels. Thus, across trials, responses of all neurons are strongly correlated and the trial-to-trial fluctuations tend to be of similar size for each neuron for a given stimulus. By contrast, the variability for any given neuron varies more strongly across stimuli. 12 Figure 3: Response-covariance relations depend on origin of correlations. A-C: Average population response (cid:104)r(cid:105) versus average (co-)variances in recurrent network (top), feed-forward net- work (middle) and gain-fluctuation model (bottom). Each dot corresponds to a random stimulus, blue dashed lines indicate analytic results, Eqs. (28)-(32). D-F: Dependence of normalized variabil- ity projected on mean direction on response direction for the three scenarios. In D,E colors indicate √ results for different networks, in F size of markers indicates strength of gain fluctuations. Square cN . G-I: Same for the normalized variability markers on the left indicate numerical value of projected on diagonal direction. See Appendix for further details and the numerical parameters. 4.3 Relation between signal correlation and noise correlation in recurrent network models In the simple scenario we analyze here, average responses and covariances are related because they both depend on the network architecture. By expressing signal correlations, cS ij, and the noise correlations, cN ij , in terms of network and input parameters, we can derive a relation between these two sets of quantities (Fig. 4). For a random transfer matrix, B, all neuron pairs are statistically equivalent, and the strength of correlations can be characterized by the average of the pairwise signal and noise correlations. In our recurrent network model, the network architecture affects noise and signal correlations through an effective quantity, 13 namely the signal-to-noise ratio of the elements of the transfer matrix, ρ = var(B)/(cid:104)B(cid:105)2: , cS ≡ (cid:104)cS ij(cid:105)i(cid:54)=j ≈ 1 1 + ρ cN ≡ (cid:104)cN ij(cid:105)i(cid:54)=j ≈ 1 + (N − 1)cin 1 + ρ + (N − 1)cin (33) (34) (see Appendix 8.3.4, Eqs. (70) and (77)). Here, cin denotes the strength of signal correlations in the input across stimuli. Both signal and noise correlations are larger in networks with more homogeneous entries (smaller ρ). We already showed in Section 4.1 that noise covariances, (cid:104)Cij(cid:105)i(cid:54)=j, depend on the average strength of the effective connections, while the variances, (cid:104)Cii(cid:105)i, depend on var(B) as well. If the average input for pairs of neurons across stimuli is uncorrelated (no input signal correlations, cin = 0), noise and signal correlations are identical on average. However, due to a prefactor which grows with system size (N − 1 in Eq. (34)), even weak input signal correlations are strongly amplified and can yield a large effect. Because noise in the population response to a given stimulus is produced internally in the process of spike generation, noise covariances are unaffected by this mechanism, so that strong signal correlations can coexist with weaker noise correlations. These relations depend on network parameters and hold for the average correlations (across pairs in a network). The pairwise noise and signal correlation coefficients, cN ij and c2 ij, can vary widely, but numerical calculations indicate that they are correlated in a given network (across pairs) as well. Figure 4: Relation between signal and noise correlations in the recurrent network model. A: Dependence of average signal correlations, cS, (continuous lines) and average noise correlations, cN , (dashed lines) on input correlation and and network properties, Eqs. (33), (34). Low variability of the transfer matrix elements, ρ, increases cN and cS. Input signal correlations affect only cS. B: Same data, average signal versus noise correlation. C: Scatter plot of all pairwise signal versus pairwise noise correlations (dots) in five network realizations for cin = 0.05. Circles indicate network average across pairs, orange dotted line corresponds to analytic expressions shown in B for cN and cS. 4.4 Stimulus discrimination in a recurrent network model Since the network architecture affects both signal and noise correlation, it is natural to ask how the discriminability of stimuli is affected in turn, as it depends on both quantities. Indeed, noise correlations can affect the coding properties of a population appreciably [19]. Noise 14 correlations are referred to as favorable for the discrimination of a pair of stimuli when the shapes of the two correlated response distributions (corresponding to the two stimuli) is such that variability occurs predominantly in a direction orthogonal to the informative direction (Fig. 5). The relevance of correlations can be quantified by comparing the discriminability of stimulus pairs in the case of the full population output and in the case in which noise correlations are removed by shuffling trials. In Fig. 5, we visualize, for a recurrent network model, the interplay of the two effects which affect the influence of correlations: on the one hand, how likely it is that the influence of correlations be favorable for a pair of stimuli, and, on the other hand, how strong the effect of correlations is for a given pair. For the first question, we utilize the fact that the noise is predominantly along the diagonal direction. Consequently, the effect of correlations is favor- able, if average responses differ in a direction orthogonal to the diagonal. The distribution of the angle between these two directions will therefore reflect the likelihood of beneficial correlations. To confirm this picture, and to examine the second question, we plot how the effect of shuffling on the signal-to-noise ratio, S, depends on this angle (see Eq. (21) for the definition of S). The network parameters, in this case the variability ρ of the elements of the effective connections, affect both the distribution of angles and the effects on S - the distribution of angles via the distribution of average responses, or the signal correlations, and S by the strength of the noise correlations. However, it turns out that the effect of shuffling, averaged across all pairs, is the same in each network. This is due to the fact that signal correlations are on average as strong as noise correlations in all of these networks. In Section 4.8, we analyze a simplified network model to better understand some of these effects. 15 Figure 5: Covariances affect stimulus discriminability A: Sketch of response distributions for two neurons and two stimuli (blue ellipses). Neural responses are correlated for both stimuli, so that main axis of variability is parallel to linear separator (orange line). B: Larger ensemble of stimuli for correlated neurons (large variance along diagonal directions). The effect of correlations is favorable for stimulus pairs where difference in means is orthogonal to diagonal (red arrow), or unfavorable when difference in means tends to be parallel to diagonal (green arrow). C: Distribution of cosines of angle between diagonal and difference in means across stimulus pairs in recurrent networks with different parameter ρ. In networks with low ρ, unfavorable angles are more frequent. D: Ratio between discriminability for correlated and shuffled distributions, dots correspond to stimulus pairs. Lower values indicate that correlations are more beneficial. Dashed horizontal lines indicate average across all stimulus pairs. 4.5 Variability and correlations in populations of mouse cor- tical auditory neurons We use the theoretical results presented in the previous sections to analyze the responses of a population of neurons to different stimuli. The data set contains the firing rates, collected during a certain time interval, in response to the presentation of different sound stimuli in a number of trials (see Section 3.1 and [32]). We will compare the properties of covariances (across trials) and average responses with the predictions of our different models. Based on the models, we can then evaluate the effect of network generated variability on stimulus discrimination. The trial-to-trial variability in single-neuron output is large (supra-Poisson, Fig. 6). Neu- rons with larger average response exhibit also a larger variability in their responses. This tendency is observable across different neurons responding to a given stimulus, for individual neurons across stimuli and for the average population response across stimuli. The pairwise correlations in trial-to-trial variability are generally high and pairs with high noise corre- lations tend to have strong signal correlations. In other words, because signal correlations measure the similarity of average responses across different stimuli, neurons with similar tun- ing properties also have similar trial-to-trial variability. Additionally, in populations with strong average signal correlations, noise correlations are strong as well. These relations be- 16 tween noise and signal correlations can be reproduced in the recurrent network model (Fig. 4). However, similar results may be possible also in alternative scenarios for the generation of correlated variability. We compare the different models in the next section. Figure 6: Variability and correlation in mouse auditory cortex. A: Average response versus response variance in a single population. Colors correspond to 10 randomly chosen stim- uli. Dots correspond to single neurons, circles to averaged values across population. Dashed line indicates identity. Inset: All neurons for all stimuli. B: Scatter plot of signal correlation coeffi- cient (across stimuli) vs. noise correlation coefficient (averaged across stimuli) for different neural populations measured in the same animal, marginal histograms at top and bottom. Signal and noise correlations are correlated across pairs. Correlations are high in general, but the amount of signal and noise correlations varies strongly across populations. Circles denote average across pairs in each populations. Inset: Black circles: Average signal and noise correlations in all measured populations. Grey dots: individual pairs, 5 % randomly chosen from all experiments. 4.6 Analysis of population response variability in mouse au- ditory cortex Motivated by our investigation of network models, we first examine the relation between the shape of the response distributions for each stimulus and the pattern of average responses for the entire set of stimuli. In Fig. 7, we display the dependence of the normalized standard deviation projected on the direction of the mean response, σµ/σall, and the normalized stan- dard deviation projected on the diagonal direction, σd/σall, on the angle between average response and diagonal, through cos(d, r) (see Section 3.2 for details and definitions). We observe a strong dependence on cos(d, r) of the variance projected on the direction of the mean response, but not of the variance projected on the diagonal direction. The dependence of σµ/σall on the stimulus is much stronger in nearly all the measured populations. This behavior is consistent with a network model, either feed-forward or recurrent, but not with a model with shared gain fluctuations, where a large part of the variance is consistently in the direction of the mean response (see Section 4.2.) Note, however, that in the data firing rates are measured relative to the spontaneous activity; this is reflected in the presence of negative values in the stimulus evoked activity. Such an offset could affect our comparison because we do not know the true value of the mean response. To account for this possibil- ity, we searched for the best possible offset, assuming the model of gain fluctuations for the 17 stimulus-dependent covariances. We then corrected the firing rates by this offset, and eval- uated the stimulus dependence of σµ and σd, as before (see Appendix 8.3.2 for details), but found no qualitative change in the results. However, we did not test for mixed models, and it is possible that part of the correlated variability can be explained by shared fluctuations. Based on the results presented above we conclude that, of our three scenarios, feed-forward and recurrent network models are more consistent with the data. In the following, we fit the parameters of these models to the data to find which of these two provides a better fit. For this comparison, we analyze the dependence of the average variances, (cid:104)Cij(cid:105)i(cid:54)=j, and covariances, (cid:104)Cii(cid:105)i, on the population-averaged response, (cid:104)r(cid:105)(s). In the recurrent network (Eqs. (27) and (28)), both variances and covariances increase linearly with (cid:104)r(cid:105) across stimuli, while in the feed-forward network (Eq. (32)), mean covariances are not expected to depend strongly on (cid:104)r(cid:105). To examine whether the experimental data is consistent with the feed-forward model, we fitted parameters to the activity of each experimentally observed population (see Appendix 8.3.5), and generated a surrogate data set of responses and covariance matrices, with a match- ing number of neurons and stimuli. In particular, we obtained values for the variability of the input ensemble, ρE = var(rext)/(cid:104)rext(cid:105)2, and of the network elements, ρ = var(F )/(cid:104)F(cid:105)2. It turns out that ρE, ρ > 1: for both input and network elements the variance is much larger than the mean (see Fig. 8A-B). The reason is the high variability of average responses across stimuli. With these parameters, the statistics of the distribution of average responses are well reproduced (Fig. 8C). As we discussed in Section 4.2, for a high value of ρE the feed-forward model does not predict a strong relation between (cid:104)Cij(cid:105)i(cid:54)=j and (cid:104)r(cid:105)(s). We quantify this re- lation by the ratio of slope and intercept from a linear fit to the data, which indicates how strongly the average covariances increase for stimuli that evoke strong population responses. In the feed-forward scenario, this behavior does not obtain: the slopes of the linear fits in the model relative to the intercepts are too low in comparison to what is observed in the experi- ment (Fig. 8D, E). The feed-forward model thus cannot reproduce both the large variability of neural responses across stimuli and the increase of the average covariance with the average response. In a recurrent model, this increase is expected, because noise generated by the neurons is propagated through the network: the covariances are proportional to the average response (see Eq. (28)). Moreover, it predicts that the ratios of intercept and slope in the linear behaviors of variances and covariances are the same (it indicates the variance of the external input). Indeed, linear fits to these relationships reveal approximately consistent ratios of intercept and slope. The estimated parameters are summarized in Fig. 8G and H. The external noise, estimated from the ratio intercept/slope, turns out to be of the same magnitude as the average rate. Interpreted in terms of the model, the noise resulting from Poisson spike generation thus contributes as much to neuron variance as the external input. This combined variability is propagated through the network, and, on average, multiplied by a factor N(cid:104)B(cid:105)2 > 1, corresponding to the slope in Eq. (28). Because this factor is larger than one, both average response and noise are amplified by the recurrent connections. In summary, we find that a model of recurrently connected neurons with random effec- tive connections captures the observed activity statistics, in particular the relations between average response and average covariances as well as the consistent direction of population fluctuations across stimuli. Both purely feed-forward networks and gain fluctuation models are not consistent with all of these observations. We note, however, that our conclusions are based on a relatively small data set. A larger number of trials and measurements of absolute values of firing rates would be desirable for a more stringent test. 18 Figure 7: Dependence of noise distribution orientation on average response in data. A, B: Typical example, 6 neural populations recorded in the same animal. Relative variances projected on mean and diagonal direction, versus cosine of angle between mean response and diagonal. Each marker corresponds to a different stimulus. Different markers/colors denote different populations. Squares to the left side indicate(cid:112)(cid:104)Cij(cid:105)i(cid:54)=j/(cid:104)Cii(cid:105)i. Solid lines: linear fit. C: Slopes from linear fits as in A/B from σµ/σall vs. slope from σd/σall for all measured populations. Circles correspond to slopes for positive, squares to negative cos(r, d). Colors in all panels indicate value of cN in populations. 19 Figure 8: Scaling of covariances with average response in experiment and model. A, B: Distribution of parameters ρ and ρE (variability of network connections and stimulus input) estimated from all experiments. C: Histograms of cos(d, r(s)) from mean responses across stimuli, for a selection of populations measured in one animal. Filled histograms: experiment. Solid lines: Model results for 500 randomly generated stimuli and one network with random effective connec- tions, parameters inferred from data. D: Population averaged response versus population averaged variances and covariances for all stimuli (full circles) in one experiment and from corresponding feed-forward model (empty circles). Dashed lines represent linear fits. E: Scatter plot of ratio slope/intercept from linear fit in all experiments versus corresponding value in the feed-forward model. Orange circle indicates population used in panel D. F: The ratio intercept/slope is, across experiments (dots), consistent for covariances and variances. G: Distribution of estimated slopes for average variances. H: Distribution of estimated (cid:104)Vext(cid:105) (intercept/slope from fit to variances) in comparison to average and minimum firing rate across experiments. 4.7 Influence of noise correlations on stimulus discrimination We quantify the influence of noise correlation on stimulus discrimination by the ratio Sshuffled/Soriginal. S is calculated for the data set before and after shuffling trials. It is defined in Eq. (21) and denotes, for a pair of stimuli, the difference in average response divided by the standard de- viation of the responses, both projected on the most discriminative direction. Larger values indicate that stimuli are easier to discriminate. If Sshuffled/Soriginal is larger than one, then 20 removing correlations by shuffling trials improves stimulus discrimination. Across pairs of stimuli, this signal-to-noise ratio varies strongly (Fig. 9). On average, stimuli are slightly easier to discriminate in the shuffled data, so that noise correlations are weakly unfavorable to the encoding of the stimulus set used in the experiments. A closer analysis of the response distributions reveals that, to a large degree, the effect of shuffling can be explained by the relative locations of the two mean responses and the diagonal, as measured by cos(d, r(s1)−r(s2)) (Fig. 5). This is because, due to the noise correlations, the main direction of variability is along the diagonal. The overall effect of shuffling on stimulus discrimination depends on the relation between noise and signal correlations: stronger signal correlations lead to a stronger dominance of angles that are unfavorable if noise is aligned along the diagonal direction, and in this case shuffling correlations away will benefit stimulus discrimination. Figure 9: Effect of correlations on stimulus discrimination. A: Signal-to-noise ratio S, dots correspond to pairs of stimuli. Value from observed covariance matrix versus the one based on shuffled trials, for different populations in the same animal. B: Same data, distribution of S(shuffled)/S(original). Broad distribution, but mean value greater one for all populations (short solid lines) indicates that shuffling increases information on average. C: Distribution of cosine of angle between diagonal and average response difference across stimulus pairs for neural popula- tions. Large cosines/small angles are most frequent. D: Scatter plot of effect of correlations on discrimination. Small dots correspond to stimulus pairs. Connected large dots indicate average value in a bin centered at the corresponding location on the x-axes. E: The distribution of mean values across all measured populations shows that correlations are on average unfavorable for al- most all populations. F: Average effect of correlations is stronger for increasing signal correlations. 4.8 Intuition from a two-population model The relation between network properties and the statistics of variability, on the one hand, and the relation between the statistics of variability and the encoding of stimuli, on the other hand, have been studied widely. Above, we describe the structures of correlations that emerge from different network architecture. We also analyze the structure of correlations in data from mouse cortex, and we relate them to stimulus encoding. Here, we study a highly idealized recurrent cortical network in order to develop an understanding of the relation between network characteristics, response statistics, and stimulus encoding. 21 To understand the effect of recurrent connections on the discriminability of a pair of stimuli, we consider two populations of excitatory neurons. The neurons in each of these populations prefer one of the stimuli, and we want to understand the effect of connections between these differently tuned neurons. The connectivity is regular, in the sense that each neuron projects to a fixed number of neurons in its own population and to a fixed number of neurons in the other population. Neurons in each population also receive the same external input. We start by considering two stimuli, s1,2, where for each stimulus one of the populations receives a stronger input. To discriminate between these stimuli, it is sufficient to consider the network response at the population level, that is, the sum of the response of neurons in each of the two populations. This follows from the regularity of the connections. The distribution of responses across trials can thus be described by a two-dimensional vector, R, of the average population response and a covariance matrix, Σ, of the variability. The sum of the couplings within a population, Γs, and across populations, Γc, determine the response via a symmetric 2 × 2 coupling matrix, Γ. Here, we will manipulate the entries of Γ, rather than the effective population coupling matrix, P = (I − Γ)−1 (see Fig. 10 and Methods for further details of the model). We start by examining how cross-coupling between the populations can induce beneficial correlations for stimulus discrimination, by considering the effect of connections between the populations on the covariances. If we introduce more connections across populations, the cross-coupling Γc increases and the response distributions evoked by the two stimuli change (Fig. 10). The responses of the two populations are amplified, but also more similar - the difference in population response, R(s1) − R(s2), decreases. Simultaneously, the correlation between the population responses increases, rendering the distributions more elongated. To evaluate the effect of shuffling trials, we need to consider the covariances between individual neuron pairs. As we show in detail in Appendix 8.4, covariances reduce the overlap of the distributions if Γc > Γs. Then, covariances between neurons belonging to different populations are larger, on average, than for neurons of the same sub-population, and the signal-to-noise ratio, S, becomes smaller if covariances are shuffled away (Fig. 10C, D). While it thus appears that strong coupling between differently tuned neurons induces beneficial pairwise correlations this is not always the case. To see this, we compare networks with different parameters, and disentangle the effects of cross-connections on average responses, variances, and covariances. Increasing the coupling between populations decreases S both in the shuffled and in the original cases (Fig. 10E), but this suppression is stronger when covariances are shuffled. The suppression can be traced back to the amount of noise produced through spike generation. Recurrent amplification increases firing rates, and the variance of Poisson neurons increases with rate. If the amount of noise remained constant, S would not depend on Γc: larger coupling induces stronger correlations which reshape noise along a direction irrelevant for discrimination, but at the same time average responses approach each other. These two effects precisely cancel each other, ultimately because a linear transformation of responses by the network cannot reduce noise. Above, we only considered a pair of stimuli for which beneficial correlations are realized in our network. We now evaluate the effect of cross-connections on a more general stimulus en- semble. We define it via the input vectors for the two populations, (1+∆ cos(φ), 1+∆ sin(φ))T for φ ∈ (0, 2π) and a constant ∆ (see Fig. 10G and Appendix 8.4.1). For this ensemble we see that inevitably correlations will be unfavorable for certain stimulus pairs, depending on the stimulus dimension in which their inputs differ. The averaged effect of shuffling across this two-dimensional stimulus ensemble depends not only on the noise correlations, but also on the distribution of the input, which affects the distribution of the average responses, i.e. the signal correlations. 22 If excitatory recurrent networks amplify internally generated noise, are feed-forward ar- chitectures generally more efficient? In the following, we compare the performances in the two scenarios. As we noted before, the difference between the two scenarios is that the noise produced by spike generation in the recurrent network is fed back into the network, and therefore correlated, while it is independent in the feed-forward network. The measure of performance we used so far is the discriminability of pairs of stimuli. We want to compare the effect on information of a recurrent network, with a transfer matrix P defined by the con- nectivity parameters, to a feed-forward network with identical transfer matrix, for changes of a high-dimensional input in arbitrary directions. A quantity that measures the informa- tion of a network response statistics in the case of a continuous stimulus is the linear Fisher information matrix. In the two-population model, it is given by in the recurrent case and by IR = (D[R] + Σext)−1 IF = PT (PΣextPT + D[R])−1P (35) (36) in the feed-forward case. Here, Σext denotes the variance of the external input to the two populations (see Appendix 8.4.3 for details). As figures of merit for the coding of high- dimensional stimuli, we use the traces of IR and IF, for the recurrent and feed-forward cases respectively. In the recurrent network, this quantity depends only on the firing rates. These, in turn, depend on the average strength of the effective connections, (cid:104)P(cid:105) (Fig. 10H): the larger (cid:104)P(cid:105), and therefore the response, the more noise is generated, and the more information is lost. In a corresponding feed-forward network, information depends also on the correlations. However, because the network-generated noise is uncorrelated, increasing firing rates effec- tively improves the signal-to-noise ratio, and therefore increases the information. Whether a feed-forward or a recurrent architecture is preferable thus depends on the output firing rate. From this model, it is not clear how firing rates should be chosen, but additional constraints, for example imposed by energy consumption are conceivable. A simple model with two coupled, uniform populations illustrates how variations in the network parameters can affect stimulus discriminability, through modulations in the average responses and variability. Broadly speaking, stimulus discriminability depends on the inter- play between noise creation, average response, and noise correlation, and the relative behavior of these quantities depends on the network architecture. 23 Figure 10: Model of two coupled excitatory populations. A: Circuit model. Neurons are coupled with excitatory connections within and across two populations. For each stimulus, one population receives stronger input. Population responses are described by macroscopic vari- ables. B: Responses to two stimuli (colored dots on dotted red line) are amplified for stronger cross-coupling (brighter colors). Simultaneously, response fluctuations are more correlated. If in- ternally generated noise does not increase with rate, variability is smaller (dashed ellipses). C and D: Distribution of pairwise covariances within (dark bars) and across populations (light bars), vertical lines indicate average. If cross-coupling is weak (C), average covariance is larger within populations than across, if it is strong (D), covariances between neurons of different populations are larger. E: Dependence of stimulus discriminability S on Γc. Strong couplings implies smaller discriminability and even more so for shuffled trials. F: Increasing cross-coupling reduces difference in mean response (black line). Effect on variance in the relevant direction differs for correlated populations (solid orange line) and model with independent neurons/shuffled correlations (dash- dotted orange line). G: Red dots: average response for stimulus ensemble. Ellipses: Response distributions for four stimuli for the recurrent network (blue) and a feed-forward network with identical average responses (red). Dotted/dashed lines: stimulus pairs where correlations have positive/negative influence.H: Linear Fisher information, dependence on average of population transfer matrix elements (horizontal axes) for recurrent and feed-forward scenario. Darker lines indicate larger variance of transfer matrix elements. Lines overlap in the recurrent scenario. 24 5 Discussion Despite large amounts of work on the topic, the interpretation of the observed response fluctuations in population activity, in the context of computation by neural circuits and information coding, is not elucidated [11]. In the present study, we seek to understand the influence of correlated variability on stimulus representation, based on the exploration of model networks of stochastic neurons together with data analysis. We modeled neural activity using a set of coupled Poisson processes in order to study the relations between covariances and average responses in a neural population, across different stimuli. We found that these relations differ depending on how covariances are generated, and that correlations observed in mouse auditory cortex are consistent with those generated in a recurrent network model. This model also helps us interpret the effect of the observed correlations on the discriminability of stimuli and the nature of the propagation of information through the network. 5.1 Origins of the structure of correlations in neural circuits In the examination of the impact of correlated response variability on stimulus represen- tation, it is useful to be aware of the possible range of its mechanistic origin. Spike train variability has been attributed to the dynamics in recurrent networks [1, 37]. Along with this source of variability, in stable states of asynchronous activity, recurrent network models can also produce transitions between different states of global activity [3, 38, 39]. Alternatively, external signals fed to sub-populations of neurons, such as gain fluctuations induced by up- stream areas, were pointed to as causing high variability and correlations [7, 15, 9, 8, 24]. Population-wide gain fluctuations can also reproduce dependencies between noise and signal correlations [10]. In general, it is difficult to differentiate shared input or gain fluctuations, and recurrent connections, as the origin of correlated variability, based on measured activity. Here, instead of focusing on temporal dynamics [40, 13, 41], we approach the problem by considering the relation between the network architecture and the structure of population activity statistics, when a neural population is presented with an ensemble of stimuli. Our analysis makes the assumption that noise resulting from spike generation in neurons can be modeled as a Poisson process, and, thus, depends on the firing rate. If the noise generated within the network is purely additive and independent of firing rates, it is indistinguishable from external noise that is simply filtered by the network [35]. In the Poisson model, however, the interplay of internally generated noise and variability due to external signals imprints its signature on the dependence of covariances on firing rates. 5.2 Influence of correlations on stimulus discrimination in neu- ral populations We use a recurrent network with random effective connections to estimate the influence of the variability generated by the network dynamics on stimulus representation. Based on the paradigm that response variability amounts to noise that downstream neurons have to cope with, a number of authors have argued that noise correlations can benefit information coding [19, 21, 23, 22, 24, 25]. Furthermore, recent studies have examined specifically the impact of correlations generated by recurrent networks. Reference [30] stresses that it is the structure of the noise, more than its magnitude, which determines the role it plays for information coding. Harmful structures of noise correlation, causing information to saturate in large neural populations, are the consequence of the amount of irreducible external noise and potentially sub-optimal processing [31], but these structures are difficult to pinpoint 25 in detail in measured activity [30]. By contrast, it was demonstrated that noise internally generated in a network via spike generation can be averaged out in feed-forward or recurrent networks, depending on the connectivity in the network and the redundancy of the population code [27]. In this work, the goal was to start from the observed properties of the activity, and to interpret mean activity and noise correlations based on outcomes of model network dynamics. Variations of network properties, such as the architecture of connections, change both the mean activity and noise correlations, both of which influence the accuracy of the neural code. Consequently, it is useful to consider both simultaneously to understand network effects. The effect of recurrent dynamics on input information depends mostly on the amount of internally generated noise. Our toy model, as well as Ref. [27], argue that recurrent amplification suppresses information. The reason is essentially that, for Poisson neurons, larger firing rates imply a higher variance. The stronger the recurrent connections, the more the average response is amplified, and the more noise is produced internally. Because this noise is fed back into the network and thus amplified furthermore, the variance is increased to the point that the signal-to-noise ratio decreases. We also show that under simplifying assumptions, including a random effective network architecture and a large, random stimulus ensemble, the amplification factor of the network and the amount of noise can be estimated from measured responses. Limitations of our analysis result, in particular, from the assumption of linearly interacting Poisson processes. Nonlinear transfer functions can affect the relation between rates and correlation in individual pairs [42, 11]. Using Poisson neurons also implies that the generation of noise in the network is attributed to individual neurons. The tractability of the model, however, allows us to obtain relations between observable statistics, like noise and signal correlations, which can be used as a starting point to interpret correlated variability in terms of its mechanistic origins. 5.3 Variability in mouse auditory cortex We tested the consistency of different scenarios involving interacting Poisson neurons with measured neural population responses, and we interpreted the observed variability in terms of a recurrent network model. The measured population activity is given as spike counts in 200 ms bins, so that correlations are defined over a relatively slow timescale and, therefore, do not take into account fast temporal structure (as noted in, e.g., Ref. [43]). Both experimental variability and correlations are high [44], so that the effects of correlations are potentially strong and amenable to an analysis such as ours. Although variability and correlations are often high in anesthetized animals [9], noise and signal correlations were somewhat weaker in comparable experiments with anesthetized animals [45]. It is conceivable that part of the correlations are due to experimental artifacts from the calcium imaging technique used in the experiments (like scattering of fluorescent light by the neuropil), but intracellular recordings of rates were consistent with calcium recordings [32]. We model the set of stimuli as a generic, high-dimensional ensemble, and the network connections at the level of a transfer matrix. When analyzing the discriminability of pairs of stimuli, this enables a broader analysis than one relying on a single stimulus dimension or av- erage correlation. The recurrent network model explains not only the scaling of variances and covariances with firing rates, but also the relation between signal and noise correlations, and the effect of shuffling away covariances on stimulus discrimination, in a natural way. Differ- ences among populations of neurons can be attributed to changes in the effective connectivity within populations. Recent studies have explained correlated variability in terms of "external" gain fluctu- 26 ations acting on a (sub-)population [7, 15, 9, 8, 24]. We showed that recurrent networks of strongly connected neurons can exhibit population-wide fluctuations as well. Estimated model parameters indicate an amplification of input signals by excitatory recurrent con- nections. This suggests that the noise that can be traced back to spike generation in the population - noise harmful to coding - is comparable in magnitude to the external input noise. Our data analysis suffers from an unknown offset in the firing rates, a relatively small number of trials, and the low temporal resolution. We therefore only compared different pro- totypical scenarios without trying to evaluate the outcome of a mixed model. In experimental populations of neurons, the statistics of activity can obviously result from a combination of both feed-forward and recurrent processes. 6 Acknowledgments We thank Brice Bathellier for providing the experimental data set as well as commenting on it, and for discussions. This work was supported by the CNRS through UMR 8559 and the SNSF Sinergia Project CRSII3 141801. 7 Author Contributions Conceived and designed the experiments: VP RS. Performed the experiments: VP. Analyzed the data: VP. Contributed reagents/materials/analysis tools: VP RS. Wrote the paper: VP RS. 27 8 Appendix 8.1 Numerical simulations: details and parameters of a rate offset a = 4, so that C = (I − G)−1(cid:0)D[r] + a + D[Vext](cid:1)(I − GT )−1. The stimulus In Fig. 2, covariances were calculated for recurrent networks from Eq. (12), with the addition ensemble consisted of 200 input vectors rext of size N = 60, with entries independently chosen from a normal distribution with mean 0.2 and standard deviation 1. Rates r were calculated as in Eq. (7), and we assumed Poisson input with Vext = rext. For the networks, four connectivity matrices G of size N = 60 were generated. Their entries were chosen from a normal distribution with standard deviation 1.5/N and mean 0.8/N − 0.9/N , respectively. For the analytic predictions from Eqs. (28)-(32), the mean and variance of the transfer matrix B = (I − G)−1 were calculated numerically. For Fig. 3, for the input ensemble for the recurrent network, elements of 200 random input vectors rext of length N = 60 were chosen from a normal distribution with mean 0.4 and variance 2. Connectivity in the recurrent network was as in Fig. 2. For rates and covariances of the feed-forward scenario, we use Eqs. (15) and (16). We set F = 10 · B, and scaled the input vectors by a factor 1/10 to obtain rates and correlations of comparable size as in the recurrent network. Negative output rates were set to zero. For the gain fluctuation model, responses of the network with ρ = 1.4 were used as the ensemble of stimulus responses. From these, covariances were calculated, see Eq. (18), for which the variance of external fluctuations Vext was varied. For this figure, no rate offset was used, a = 0. The networks in Fig. 4C were identical to the ones used above. 50 stimuli rext with random entries were generated with mean 0.2 and standard deviation 1.5. Inputs for different neurons were correlated with correlation coefficient cin = 0.05 across stimuli. Rate offset was 4. In Fig. 5, networks and stimulus parameters were the same as in Fig. 4, but there were no input signal correlations, cin = 0. In Fig. 10 the network connects two populations of 100 neurons each. For the pair of stimuli, panel B, neurons in the population that was more strongly excited received an input of 1 + ∆, with ∆ = 0.2, and the others input 1. More general inputs (panel G) were defined by setting the inputs Rext = (1 + cos(φ)∆, 1 + sin(φ)∆) for φ ∈ (0, 2π). Each neuron had a fixed number of nEE postsynaptic partners within the same populations. Coupling strength between connected neurons was set to g = 0.01. We chose a population coupling within populations of Γs = 0.2, and the number of connections between neurons within populations such that Γs = nEEg. Connections across populations were chosen accordingly to realize across coupling values Γc = 0, . . . , 0.4. To visualize the effect of the increase in internally generated noise due to increased rates (panel B), we compared the covariances Σ = BpD[R](Bp)T for increasing Γc to the ones where in D[R] the rates of a network with Γc = 0 were used. In panels E and F, we calculated Soriginal by inserting the population covariances and responses in Eq. (21) for the two stimuli, for the correlated case. For the shuffled case, the full matrix covariance matrix C was generated and the off-diagonal elements set to 0, such that (cid:18)(cid:80) Σshuf f led = . (37) (cid:19) (cid:80) i∈E Cii 0 0 i∈E(cid:48) Cii For panels E-H, population level parameters in the matrices Γ or P were varied directly. We set the external noise Σext = N D[Rext]. In panel G, Γc = 0.4 and Γs = 0.2. In panel H, we fixed for the population transfer matrix P the ratio ρ = (cid:104)P(cid:105)2/var(P ) to the values ρ = 0.5, 1, 2. For each of these values, (cid:104)P(cid:105) was varied between 0.1 . . . 5, so that var(P ) was fixed. From (cid:104)P(cid:105) and var(P ), we calculated Ps and Pc, see Appendix 8.4.1. 28 8.2 Common framework for the generation of correlations Here we formulate the three model scenarios described in Sections 3.3 and 3.4 as special cases of the interacting point processes framework defined in Eq. (6). The derivation is similar as in [33, 27]. We consider an extended network of two populations which receive only constant input rfull(t) ≡ rfull. Neurons are divided into an observed population O and an unobserved external population U . The external population projects to the observed population, but does not receive feedback. The coupling matrix of integrated kernels of the full system is a block matrix of the shape Gfull = . (38) (cid:18)E 0 (cid:19) F G The coupling between nodes within the external network is described by the matrix E. If E = 0, external inputs are independent Poisson processes. The feed-forward weights to the observed network are defined by F , and by G the recurrent connections within the observed population. The input vector to the system is rfull = (r0, 0). The components of the vector r0 together with E determine the firing rates of the input nodes. The 0 represents a vector of zeros, so that there is no constant input directly to the observed population. The time dependent firing rates of neurons k ∈ U in the external input population are determined by rk(t) = ekl(τ )sl(t − τ )dτ + r0,k and the rates of neurons i ∈ O in the observed population by (cid:90) ∞ 0 (cid:88) l∈U (cid:90) ∞ (cid:88) j∈O 0 (cid:90) ∞ (cid:88) k∈U 0 ri(t) = gij(τ )sj(t − τ )dτ + fik(τ )sk(t − τ )dτ. (39) (40) (41) (42) The transfer matrix of the full system is (I − E)−1 Bfull = (I − Gfull)−1 = (I − G)−1F (I − E)−1 (cid:18) (cid:19) ≡ (cid:18) BE 0 BF BE B. (cid:19) 0 (I − G)−1 The average rates of the system are, from applying Eq. (7) to the full system, (I − Gfull)−1rfull = Bfullrfull = (BEr0, BF BEr0)T ≡ (rext, r)T , where rext = BEr0 are the rates of the input neurons and r = BF BEr0 = BF rext corre- spondingly, the rates of the observed neurons. From Eq. (12), the covariance matrix is given by Cfull = BF BED[rext]BT E B BED[rext]BT EF T BT D[r] + F BED[rext]BT . BT (43) (cid:32) BED[rext]BT E (cid:16) (cid:33) EF T(cid:17) The block CE ≡ BED[rext]BT population. The block at the lower right, E at the upper left describes the covariances of the external C ≡ B(D[r] + F CEF T )BT (44) describes how covariances in the observed network depend on the properties of the unobserved input neurons. The recurrent network scenario is obtained for an input population and an output popu- lation of equal size N , with F = I and any diagonal E. In this case, CE is diagonal, and we can identify the elements on the diagonal of CE with Vext. 29 For the feed-forward scenario set G = 0 and again identify the diagonal elements of CE with Vext. The gain fluctuation model is obtained for G = 0, and if the input population consists of a single neuron. The matrix F then corresponds to a single column vector, and C = D[F rext] + F CEF T . With r = F rext and setting Vext = CE/r2 ext (rext is a number in this case), C = D[r] + rrT Vext. 8.3 Further details on statistics in different model scenarios 8.3.1 Gain fluctuation model We derive relations between covariances and average response in the gain fluctuation model, in particular the scaling of average covariances with average rates and the orientation of the response distribution measured by the projections of the variances in different directions. From Eq. (18), pairwise covariances are directly related to average responses, Cij = δijri + rirjVext. (45) The relation between average covariance and the population averaged response, Eq. (32), follows from (cid:104)Cij(cid:105)i(cid:54)=j = 1 N (N − 1) rirjVext ≈ Vext N 2 = Vext(cid:104)r(cid:105)2, (46) rj (cid:16)(cid:88) (cid:17)(cid:16)(cid:88) ri i j (cid:17) (cid:88) i(cid:54)=j where the terms r2 i can be neglected if N is large. To measure changes of the response distribution across stimuli, we use the variance pro- d, respectively. In µ does not strongly depend on the jected in the direction of mean response and diagonal direction, σ2 the following, we give an argument that in this model σ2 stimulus, while σ2 µ and σ2 d does. Both quantities are normalized by the sum of the variances σ2 all = Cii = i Vext + ri) = r2Vext + N(cid:104)r(cid:105). (r2 √ The variance projected in the diagonal direction ¯d = (1, . . . , 1)T / i N is (cid:17) (cid:88) (cid:88) (cid:16) (cid:88) i (cid:88) (47) (48) (49) σ2 d = Cij(s) ¯di ¯dj = rirjVext + δijri /N = N Vext(cid:104)r(cid:105)2 + (cid:104)r(cid:105). ij ij To see that there is a strong dependence of σ2 cos(d, r) = ¯d¯rT = N(cid:88) i=1 d/σ2 all on stimulus direction, note that rir = √ N(cid:104)r(cid:105)/r. 1√ N For non-vanishing r, it follows from cos(d, r) = 0 that (cid:104)r(cid:105) = 0 and in this case σ2 If cos(d, r) = 1, r ∝ d and r2 = N(cid:104)r(cid:105)2, and so σ2 N Vext(cid:104)r(cid:105)2+N(cid:104)r(cid:105) . This value is of the order of one, if the ratio of Vext(cid:104)r(cid:105)2 and (cid:104)r(cid:105), which is of the order of the average noise correlation coefficient, is not very small (in comparison to one). Consequently, if noise correlations are not too small, the normalized variance projected on the diagonal direction strongly depends on the direction of the response vector. all = N Vext(cid:104)r(cid:105)2+(cid:104)r(cid:105) d/σ2 d/σ2 all = 0. The variance projected on the mean direction is σ2 µ = ¯ri¯rjCij = ¯ri¯rj(rirjVext + δijri) > Vext j /r2 = Vextr2 i r2 r2 (50) (cid:88) (cid:88) (cid:88) ij ij ij 30 Hence, σ2 r2Vext+N(cid:104)r(cid:105) , which depends only weakly on (cid:104)r(cid:105), if N(cid:104)r(cid:105) is not much bigger than Vextr2 ≥ VextN(cid:104)r(cid:105)2, that is once again if noise correlations are not too small. µ/σ2 all > Vextr2 8.3.2 Additional test for a common origin of shared variability If rates are only measured up to an unknown offset a, the variance σ2 all, which is obtained from a projection on the apparent average response r − a may not be constant across stimuli. In this case, the variation of this quantity across stimuli is not a reliable indicator to exclude a gain fluctuation model as the source of observed correlations. However, if a large part of the covariances can be explained by a single component, it should be possible to reconstruct the common origin from the stimulus dependent covariance matrices. From µ/σ2 C(s) =(cid:0)r(s) + a(cid:1)T(cid:0)r(s) + a(cid:1)Vext + D(cid:0)r(s) + a(cid:1) largest eigenvalue of C, by neglecting the contribution of D(cid:0)r(s) + a(cid:1). This approximation the vectors v(s) = r(s) + a can be obtained approximately by finding the eigenvector with the (51) can also be avoided by applying a factor analysis with a single latent component. Because a is constant across stimuli, up to measurement errors of the estimated covariances, the straight lines through the mean responses r + x¯v, for x ∈ IR and normalized directions ¯v, will intersect in the point −a. The best intersection point of multiple lines in the least square sense is given by (cid:16)(cid:88) I − ¯v(s)¯v(s)T(cid:17)−1(cid:16)(cid:88) − a = (I − ¯v(s)¯v(s)T )r(s) (cid:17) (52) s s and can be used to correct the average responses to r(cid:48)(s) = r(s) + a. We found that the analysis of σ2 d based on the corrected responses lead to no qualitative change in the stimulus dependence, indicating that a potential shared component is too weak to be identified based on the available data. µ and σ2 8.3.3 Feed-forward model Here we derive characteristic relations for the response statistics resulting in a feed-forward network. They illustrate qualitative differences between the predictions of different models and will be used to extract model parameters from the data. In the feed-forward model the mean responses and covariances are, from Eqs. (15) and (16) and allowing an offset in observed rates, r = F rext, FikFjkVext,k. (53) (54) Cij = δij(ri + a) + (cid:88) k In contrast to the recurrent network model (see below), the average covariance is not correlated to the population averaged rate (cid:104)r(cid:105) across stimuli, Eq. (30). This follows after taking the average across neurons: (cid:104)Cij(cid:105)i(cid:54)=j = 1 N (N − 1) FikFjkVext,k ≈ N(cid:104)Fik(cid:105)ik(cid:104)Fjk(cid:105)jk(cid:104)Vext,k(cid:105)k = N(cid:104)F(cid:105)2(cid:104)Vext(cid:105). (55) (cid:88) i(cid:54)=j,k Input variances Vext are assumed to be independent of the network structure and Fik indepen- dent of Fjk, which means that the strengths of connection of an external neuron to different internal neurons are independent. If (cid:104)Cij(cid:105)i(cid:54)=j is to be uncorrelated to (cid:104)r(cid:105) across stimuli, the average input variance (cid:104)Vext(cid:105) needs to be approximately uncorrelated to (cid:104)r(cid:105). This holds in our model even though rk = 31 (cid:80) 0, or the distribution of column sums(cid:80) i Fikrext,k and in each input channel the variance equals the strength of the input, Vext,k = rext,k, if either the distribution of inputs across neurons is approximately symmetric around i Fik of the feed-forward matrix (across columns) is symmetric around 0. The first case can be realized in a feed-forward network with a similar number of excitatory and inhibitory input channels. Intuitively, inhibitory inputs decrease the average output, but contribute positively to the average variance, and thus decorrelate the two quantities. Formally, one compares the random variable(cid:80) the variable(cid:80) i Fik to k rext,k. A random variable x is uncorrelated to its absolute value x if its distribution is symmetric around 0. Here, the relevant variables are the elements of the input vector rext, and their distribution is approximately symmetric around 0, if the variance var(rext) is much larger than their mean (cid:104)rext(cid:105), that is if ρE = var(rext)/(cid:104)rext(cid:105) (cid:29) 1. Consequently, if the variance of external inputs across stimuli is high, the population averaged covariances in a feed-forward network are uncorrelated to the population response. ik Fikrext,k =(cid:80) k Vext,k =(cid:80) k rk =(cid:80) (cid:80) k rext,k In the following, we motivate that the stimulus dependence of the variances projected along different directions in the feed-forward model is different from the one in the gain fluctuation model. The sum of the variances is ikVext,k ≈ N(cid:0)(cid:104)r(cid:105) + a(cid:1) + N 2(cid:104)Vext(cid:105)(cid:104)F 2(cid:105). F 2 all = N ((cid:104)r(cid:105) + a) + σ2 (56) Due to the Poisson spike generation, there is a linear contribution in (cid:104)r(cid:105) to the average variance, (cid:104)Cii(cid:105)i = σ2 all/N . This variability does not contribute to covariances, because spikes are generated independently across neurons, in contrast to the recurrent model, where the spiking or not spiking of a neuron directly influences post-synaptic firing rates. The variance projected onto the diagonal direction is d = (cid:104)r(cid:105) + a + σ2 1 N FikFjkVext,k ≈ (cid:104)r(cid:105) + a + N 2(cid:104)Vext(cid:105)(cid:104)F(cid:105)2. (57) (cid:88) i,k (cid:88) ijk d/σ2 all depends only weakly on (cid:104)r(cid:105) if the term N(cid:104)Vext(cid:105)(cid:104)F(cid:105)2 is large Consequently, the ratio σ2 against (cid:104)r(cid:105)+a, that is if the sum of covariances is larger than the sum of variances. This is the case, if correlation coefficients are larger than of the order O(1/N ). The variance projected (cid:88) onto the direction of the mean response is (cid:88) (cid:80) FikFjkVext,k ¯ri¯rj ≈ a + i r3 i r2 + (cid:104)r(cid:105)2 r2 N 3(cid:104)F(cid:105)2(cid:104)Vext(cid:105). σ2 µ = (ri + a)¯r2 i + i ijk (58) µ/σ2 all, assume again that the sum across variances in the first term a + To see that the ratio σ2 all strongly depends on the population response, in contrast to σ2 d/σ2 is not too large against the sum across covariances in the second term. The second term depends strongly on (cid:104)r(cid:105): it is 0 for cos(r, d) = 0. For cos(r, d) = 1 it becomes N 2(cid:104)F(cid:105)2(cid:104)Vext(cid:105). In this case, σ2 µ/σ2 is not too small, if ((cid:104)r(cid:105) + a)/(N Vext) is not much larger than 1 (assuming that (cid:104)F 2(cid:105)/(cid:104)F(cid:105)2 is all of order one). i r3 r2 i Finally, we calculate the average noise correlation coefficient (across stimuli and neuron (cid:80) pairs) cN = (cid:104) (cid:112)Cii(s)Cjj(s) Cij(s) (cid:105)s,i(cid:54)=j. (59) For a given stimulus, we assume that the Cij are approximately independent, so that we can write . (60) (cid:112)Cii(s)Cjj(s) Cij(s) (cid:104) (cid:105)i(cid:54)=j ≈ (cid:104)Cij(s)(cid:105)i(cid:54)=j (cid:104)Cii(s)(cid:105)i 32 Then, from Eqs. (55) and (56), cN = N(cid:104)F(cid:105)2(cid:104)Vext(cid:105) (cid:104)r(cid:105) + a + N(cid:104)Vext(cid:105)(cid:104)F 2(cid:105) . (61) Signal correlations can be calculated analogously as for the recurrent model, see below. 8.3.4 Recurrent network model The response distributions in the recurrent network are characterized by the average and covariances of responses given in Eqs. (7) and (12), and r = Brext C = BD[reff ]B, (62) (63) with reff = r+a+Vext, including an offset a. To compare this model to the alternative scenarios described in the previous sections, we calculate the relation between average (co)variances and the projected variances and r as well as the signal and noise correlations. We first derive Eqs. (27) and (28). The average covariance is (cid:104)Cij(cid:105)i(cid:54)=j = 1 N (N − 1) BikBjkreff,k ≈ N(cid:104)BikBjk(cid:105)ijk(cid:104)reff,k(cid:105)k. (64) (cid:104)Bij(cid:105)(cid:104)(cid:80) (cid:80) We assume that N is sufficiently large and that the external input characterized by Vext, rext In that case reff,k is approximately uncor- does not depend on the recurrent network. related to Bik (including the contribution of rk, because (cid:104)Bijrk(cid:105) = (cid:104)Bij l Bklrext,l(cid:105) ≈ l Bklrext,l(cid:105)). The term (cid:104)reff(cid:105) = (cid:104)r(cid:105) + a + (cid:104)Vext(cid:105) is linear in (cid:104)r(cid:105), with an uncorre- lated contribution (cid:104)Vext(cid:105), see Appendix 8.3.3. If the elements of B are pairwise independent, (cid:104)Cij(cid:105)i(cid:54)=j = N(cid:104)B(cid:105)2((cid:104)r(cid:105) + a + (cid:104)Vext(cid:105)). Under the same assumptions, the sum of variances is (cid:88) i(cid:54)=j,k (cid:88) σ2 all = ikreff,k = B2 ik(Vext,k + rk + a) ik (Vext,k + rk + a) = N 2(cid:104)B2(cid:105)((cid:104)r(cid:105) + a + (cid:104)Vext(cid:105)). Using the same argument as in the feed-forward scenario following Eq. (55), the term N 2(cid:104)B2(cid:105)(cid:104)Vext(cid:105) only contributes with an offset to the relation between (cid:104)r(cid:105) and (cid:104)Cii(cid:105)i = σall/N . (cid:104)r(cid:105) as the ones in the feed-forward model. The projected variance on the diagonal is Next, we show that the projected variances in this scenario have a similar dependence on σ2 d = 1 N BikBjkreff,k ≈ N 2(cid:104)B(cid:105)2((cid:104)r(cid:105) + a + (cid:104)Vext(cid:105)). Consequently, in the ratio σ2 all the dependence on (cid:104)r(cid:105) is canceled out. d/σ2 The projection along the mean response direction, (cid:88) ij,k σ2 µ = BikBjk rk + a + Vext,k (cid:17) ¯ri¯rj ≈ N 3(cid:104)r(cid:105)2 r2 (cid:104)B(cid:105)2((cid:104)r(cid:105) + a + (cid:104)Vext(cid:105)), 33 (cid:88) ≈ N(cid:104)B2(cid:105)(cid:88) B2 i,k k (cid:88) ij,k (cid:16) (65) (66) (67) (68) depends strongly on (cid:104)r(cid:105). Because cos(d, r) = 0 implies (cid:104)r(cid:105) = 0, σ2 for cos(d, r) = 1, with r2 = N(cid:104)r(cid:105)2 one gets σ2 all = (cid:104)B(cid:105)2/(cid:104)B2(cid:105). Apart from the raw covariances, we are interested the average noise correlation coefficient (across stimuli and neuron pairs). As for the feed-forward model, we approximate cN ≈ (cid:104)Cij(cid:105)s,i(cid:54)=j (cid:104)Cii(s)(cid:105)s,i all = 0 in this case, and . From Eqs. (65) and (66) µ/σ2 µ/σ2 cN = N(cid:104)B(cid:105)2((cid:104)r(cid:105) + a + (cid:104)Vext(cid:105)) N(cid:104)B2(cid:105)((cid:104)r(cid:105) + a + (cid:104)Vext(cid:105)) (cid:104)B(cid:105)2 (cid:104)B2(cid:105) , = and because (cid:104)B2(cid:105) = var(B) + (cid:104)B(cid:105)2 cN = 1 1 + var(B)/(cid:104)B(cid:105)2 = 1 1 + ρ . (69) (70) We see that the correlation coefficient depends on the relative variability of the network elements. The variances, resulting from the variances of the input channels, are determined by the mean of the square elements of the network. By contrast, the covariances depend on the effective weights of inputs to the neuron pairs, and hence on the square of the mean weight. Correspondingly, we consider the signal covariances CS ij(s) = cov(ri(s), rj(s))s, CS ij = cov( Bikrext,k(s), Bjlrext,l(s))s = cov(Bikrext,k(s), Bjlrext,l(s))s. (71) BikBjkcov(rext,k(s), rext,l(s))s BikBjkvar(rext,k(s))s + BikBjkcov(rext,k(s), rext,l(s))s. kl (cid:88) (cid:88) (cid:88) kl For i (cid:54)= j, one gets CS ij = = k Averaged across neurons (cid:88) kl (cid:88) k(cid:54)=l (cid:88) (cid:88) k(cid:54)=l k(cid:54)=l (72) (73) (74) (75) (76) (cid:104)CS ij(cid:105)i(cid:54)=j = N(cid:104)B(cid:105)2var(rext) + N (N − 1)(cid:104)B(cid:105)2cinvar(rext). For the signal variances, i = j, which correspond to the variance of the rates across stimuli, (cid:104)CS ii (s)(cid:105)s = = (cid:88) (cid:88) k k var(Bikrext,k) + B2 ikvar(rext,k) + cov(Bikrext,k, Bilrext,l) cov(rext,k, rext,l)BikBil. Averaged across neurons, with (cid:104)B2(cid:105) = (cid:104)B(cid:105)2 + var(B): (cid:104)CS ii (s)(cid:105)s,i = N var(rext)((cid:104)B(cid:105)2 + var(B)) + N (N − 1)(cid:104)B(cid:105)2var(rext)cin. Their ratio is (cid:104)Cij(cid:105) (cid:104)Cii(cid:105) = = N(cid:104)B(cid:105)2var(rext)(1 + (N − 1)cin) N var(rext)((cid:104)B(cid:105)2 + var(B)) + N (N − 1)(cid:104)B(cid:105)2var(rext)cin 1 + var(B)/(cid:104)B(cid:105)2 + (N − 1)cin 1 + (N − 1)cin . This results in an approximate expression for the average signal correlation coefficient ij(cid:113) cS = (cid:104) CS CS ii CS jj ij(cid:105) (cid:105)i,j ≈ (cid:104)CS (cid:104)CS ii(cid:105) = 1 + (N − 1)cin 1 + var(B)/(cid:104)B(cid:105)2 + (N − 1)cin . (77) 34 8.3.5 Estimation of network model parameters We extract parameters of the recurrent and feed-forward network models from the data, both to test the consistency of the models with the data and to interpret the observed variability. Because rates and covariance matrices were measured for many different stimuli and thus provide a large number of constraints, one approach would be to infer as much information as possible about the full connectivity matrices B or F from the data. However, due to the relatively small number of trials for each stimulus, we use a model with few parameters. The set of parameters consists of the network parameters (cid:104)B(cid:105) and var(B) ((cid:104)F(cid:105) and var(F ), respectively) as well as the parameters of the input ensemble, (cid:104)rext(cid:105), var(rext) and cin. In particular, we want to infer the ratios ρE = var(rext)/(cid:104)rext(cid:105)2 and ρ = var(B)/(cid:104)B(cid:105)2 (and correspondingly for F ). Based on these, we can generate surrogate data to test if the observed scaling of average covariances with average rates is more consistent with the recurrent or the feed-forward model, Eq. (28) or (30). The experimental data provides constraints in the form of the population and stimulus averaged rates, their variances and the noise and signal correlation coefficients. In the models, rates are given by Eqs. (7) and (15), respectively. The population averaged mean response thus is (cid:104)r(cid:105) = N(cid:104)B(cid:105)(cid:104)rext(cid:105) or (cid:104)r(cid:105) = N(cid:104)F(cid:105)(cid:104)rext(cid:105). The variance of rates across stimuli, see Eq. (75), is var(r) = N var(rext)((cid:104)B(cid:105)2 + var(B) + (N − 1)(cid:104)B(cid:105)2cin) such that the relative variance of rates var(r) (cid:104)r(cid:105)2 = var(rext) (cid:104)rext(cid:105)2 1 + ρ + (N − 1)cin N (78) (79) (80) depends on the input signal to noise ratio ρE = var(rext) elements, ρ = var(B) respectively. (cid:104)rext(cid:105)2 and the variability of the network (cid:104)B(cid:105)2 , or ρ = var(F ) (cid:104)F(cid:105)2 The estimates for the remaining parameters ρ and cin are obtained from the measured values of the ratio of covariances to variances, which correspond approximately to the average coefficients of noise and signal correlations, and (cid:104)(cid:104)Cij(cid:105)i(cid:54)=j (cid:104)Cii(cid:105)i (cid:105)s ≈ cN , (cid:104)cov(ri, rj)(cid:105)i(cid:54)=j (cid:104)var(ri)(cid:105)i ≈ cS, (81) (82) using Eqs. (70) and (77). Together with Eq. (80), these relations provide the necessary constraints for the network models. Strictly speaking, these equations are valid only for the recurrent network, while for the feed-forward model, Eq. (61) is relevant. In this case ρ is overestimated, which results in a lower bound for ρE, and thus a conservative estimate of the input variability. Under additional assumptions, we can also choose the absolute values of the parame- ters, for example (cid:104)rext(cid:105) and (cid:104)F(cid:105), such that mean rates and mean covariances correspond to experimental ones. The mean rates (78) constrain the product of the two parameters (cid:104)r(cid:105) = N(cid:104)F(cid:105)(cid:104)rext(cid:105). (83) 35 From Eq. (55) it follows that (cid:104)Cij(cid:105)i(cid:54)=j = N(cid:104)F(cid:105)2(cid:104)Vext(cid:105), (84) and we assume that Vext = rext to relate mean input and input variance. The distribution of Vext thus is a folded normal distribution, with (cid:104)rext(cid:105) =(cid:112)var(rext)(cid:112)2/πe−1/2ρx + (cid:104)rext(cid:105)(cid:16) (cid:17) 1 − 2Φ(−(cid:112)1/ρx) (with the cumulative normal Φ ), and from this one finds (cid:104)Cij(cid:105) = (cid:104)F(cid:105)(cid:104)r(cid:105)(cid:104)√ (cid:105) ρxe−1/2ρx + 1 − 2Φ(−(cid:112)1/ρx) . (85) (86) To set the absolute values of input versus network strength we made assumptions regarding the relation between input strength and input variance across trials. If we measure the strength of the dependence between mean response and mean covariance, Eq. (55) by the ratio intercept/slope of a linear fit, however, the result is independent of the absolute values of (cid:104)F(cid:105) and (cid:104)rext(cid:105), as well as a potential linear factor relating Vext and rext. 8.4 Covariances and rates on a population level in regular networks Here we derive the macroscopic Eqs. (25) and (26) for the responses of the two populations: i∈I ri, and the covariances between responses k∈K,l∈L Ckl, in the regular networks defined in Section 3.6 depend only on the coupling between populations, see below. The two populations are of size N , and weights between connected neurons are of equal strength. By definition, the regularity of the network connectivity implies that for any neuron k∈K Gkl(cid:48) is identical, and we can define the population the average response of population I, RI ≡(cid:80) of populations L, K, ΣKL ≡ (cid:80) l(cid:48) ∈ from population L, the sum (cid:80) ΓKL ≡ (cid:88) coupling matrix, Γ, by Gkl(cid:48) = 1 N Gkl = nKLg, (87) (cid:88) k∈K k∈K,l∈L where g is the weight of the connections. We want to express R and Σ in terms of Γ. On the neuron level, rates, r, and covariances, C, depend on the transfer matrix, B = (I− G)−1. We will define an analogous population transfer matrix P that depends only on Γ and show that R and Σ can be written in terms of P. We define P ≡ (I − Γ)−1 = I + Γ + Γ2 + ··· = Γm. To relate P to the microscopic quantities r and C, we will use that (cid:88) k∈K,l∈L PKL = 1 N (cid:88) k∈K Bkl = Bkl(cid:48) ∞(cid:88) m=0 for any l(cid:48) ∈ L, in complete analogy to Γ. To see this, we note the corresponding relation for all of the individual terms m (88) (89) (90) [Γm]KL = 1 N [Gm]kl. k∈K,l∈L (cid:88) 36 (cid:88) I As an example consider m = 2. Using the regularity, we see that [G2]kl = GkiGil = ΓKI Gil = N k∈K,l∈L l∈L,k∈K,i∈I,I l∈L,i∈I,I For the rate of population I, RI , we then find with (89) that (cid:88) RI ≡(cid:88) i∈I (cid:88) (cid:88) (cid:88) (cid:88) ri = Bikrext,k = Bikrext,k = i∈I k∈K K i∈I,k∈K K because rext,k = Rext,K for any k ∈ K, and thus R = PN Rext = (I − Γ)−1N Rext, ΓKI ΓIL = [Γ2]KLN. (91) (cid:88) K PIKN Rext,K, (92) (cid:88) (cid:88) (cid:88) (cid:88) (cid:88) I which is Eq. (25). Similarly, for the covariances of the population responses, Σ: Ckl = BkiBliri = PKI Bliri k∈K,l∈L k∈K,l∈L,i∈I (cid:88) (cid:88) (cid:88) l∈L I i∈I ΣKL ≡ (cid:88) (cid:88) (cid:88) I I,i∈I = which corresponds to Eq. (26), PKI PLI ri = PKI PLI RI , Σ = (I − Γ)−1D[R](I − Γ)−1, (with the diagonal matrix D[R] with D[R]IJ = δIJ RI ). 8.4.1 Effect of cross-coupling on correlations on a population level In this section, we evaluate how noise and signal correlations in the population activity arise from direct and effective coupling. The strength of the cross-coupling between populations, Γc, determines via the effective cross-coupling, Pc, of the population transfer matrix, (cid:18)1 − Γs Γc (cid:19) (cid:18)Ps Pc (cid:19) Pc Ps Γc 1 − Γs ≡ , (97) P = (I − Γ)−1 = 1 (1 − Γs)2 − Γ2 c the population rates and covariances. The network is unstable if the eigenvalues of Γ are larger than one, which corresponds to the constraint Γs + Γc < 1, and therefore Pc < Ps. On the population level, the effects of cross-coupling between differently tuned neurons on correlations can be expressed in terms of the variance and the mean of the elements of P, similar as for the random effective network model. To see this, we consider an ensemble of (two-component) input vectors Rext(s) =(cid:0)rext(s), rext(cid:48)(s)(cid:1)T with external input components rext, rext(cid:48) chosen with mean and variance µrext, σ2 rext, independently across stimuli s. The signal covariance matrix of mean responses R = PRext across the stimulus ensemble ij = cov(Ri, Rj) which is is ΣS (cid:18)P 2 (cid:19) ΣS = σ2 rext s + P 2 c 2PsPc 2PsPc P 2 s + P 2 c The average signal correlation for the population activity is (93) (94) (95) (96) (98) (cid:104)cS(cid:105) = 1 4 (cid:88) ij ij(cid:113) ΣS ΣS (cid:104)P(cid:105)2 iiΣS jj = (cid:104)P(cid:105)2 + var(P ) = 1 + 2PsPc s + P 2 P 2 c = (Ps + Pc)2 (Ps + Pc)2 + (Ps − Pc)2 (cid:104)P(cid:105)2 (cid:104)P 2(cid:105) = = 1 1 + var(P )/(cid:104)P(cid:105)2 , (99) 37 using (cid:104)P(cid:105) = (Ps + Pc)/2 and 2var(P ) = (Ps − (Ps + Pc)/2)2 + (Pc − (Ps + Pc)/2)2 = (Ps − Pc)2/2. The normalized variance of the elements of P thus determines the strength of signal correlations, and one can interpret ρ = var(P )/(cid:104)P(cid:105)2 as a measure for coupling strength. A similar relation holds for the average noise correlations, if we average first across stimuli, then across neurons: the average of the noise covariance matrix across stimuli is (cid:104)Σ(s)(cid:105)s = (cid:104)P R(s)P T(cid:105) = P(cid:104)R(cid:105)P T . Because (cid:104)Rext(cid:105) = (cid:104)R(cid:48) ext(cid:105), the average rates of both populations are identical, and the average across population pairs can be evaluated just as for the signal correlations. 8.4.2 Condition for favorable correlations Here, we show that strong connections between populations can induce correlations that are beneficial for stimulus discrimination, when compared to shuffled trials. Shuffling correlations is favorable for stimulus discrimination if the covariances within each populations are stronger than the ones across populations. To see this, assume that the responses of the excitatory populations for two stimuli are R(s1) = (R0 + ∆R, R0) and R(s2) = (R0, R0 + ∆R). We also assume that the population covariance matrix is stimulus independent, Σ(s1) = Σ(s2) = (cid:18) ΣEE ΣE(cid:48)E (cid:19) (cid:88) i∈E w = ΣEE − ΣE(cid:48)E = σ2 and that ΣEE = ΣE(cid:48)E(cid:48). Calculating the most discriminative direction w ΣE(cid:48)E ΣE(cid:48)E(cid:48) (using Eq. (20) on the macroscopic variables) yields an eigenvector of Σ , (1,−1), and the variance in that direction corresponds to the smaller eigenvalue of Σ, (cid:88) the case if the sum of covariances across populations, (cid:80) within a population, (cid:80) Covariances can be considered helpful, if this eigenvalue is smaller when compared to a shuffled version where all covariances are set to 0, such that σ2 i∈E Cii. This is exactly i∈E,k∈E(cid:48) Cik, is larger than the one i(cid:54)=j∈E Cij. Although ΣEE ≥ ΣE(cid:48)E, it is possible that (cid:104)Cij(cid:105)i(cid:54)=j∈E < Cij − (cid:88) w(cid:48) =(cid:80) i∈E,k∈E(cid:48) (cid:104)Cij(cid:105)i∈E,j∈E(cid:48), because of the contribution of the variances Cii to ΣEE. noise in the relevant direction. For simplicity, let us assume that Rext = Rext(cid:48) ≡ R. Then We want to find a condition for which such a removal of covariances by shuffling decreases i(cid:54)=j∈E Cii + Cik. (cid:18)P 2 Σ = R s + P 2 c 2PsPc 2PsPc P 2 s + P 2 c (cid:19) (100) (101) (102) The average pairwise across-covariance is Cacross = ΣE(cid:48)E/N 2. Approximately, the average neuron variance differs from the average within-covariance only by the contribution of the rate: (cid:104)Cii(cid:105) ≈ R/N + Cwithin. From this we get for the population variance ΣEE = R(P 2 s + c ) = N(cid:104)Cii(cid:105) + N (N − 1)Cwithin = R + N 2Cwithin. For favorable correlations, we need P 2 Cacross > Cwithin, 2PsPcR N 2 > R(P 2 s + P 2 N 2 c ) − R and since Ps > Pc 1 > (Ps − Pc)2 ⇔ 1 > Ps − Pc Using Eq. (97) for the entries of P , this means that 1 − Γs − Γc (1 − Γs)2 − Γ2 ⇔ Γs(1 − Γs) > Γc(1 − Γc). 1 > (103) The function Γ(1 − Γ) is a parabola through the points (0,0) and (0,1), with its minimum at (1/2,-1/4). We have Γs,Γc < 1, so both sides are < 0 and Γc < 1 − Γs. Thus, the condition can be fulfilled if Γs < 1/2 and within-coupling is smaller than cross-coupling, Γs < Γc < 1 − Γs. c 38 8.4.3 Linear Fisher information We use the signal-to-noise ratio, S, to measure how well discrete pairs of stimuli in a given ensemble can be discriminated, Section 3.5. In a model network, we have access to all possible input dimensions, and we can use an alternative measure that combines these possible stimulus dimensions. We will use this measure to compare the effects of different network scenarios on the representation of general stimuli. If a high-dimensional stimulus varies continuously, the information in the response distribution about the stimulus can be measured by the Fisher information matrix [16]. An approximate value, the 'linear Fisher information', is obtained by neglecting the information in the stimulus dependence of the covariance matrix, and the entries of the linear Fisher Information matrix are defined by Imn = ∂mRT Σ−1∂nR, (104) for the population covariance matrix Σ and population response R. Here, ∂mR is the deriva- tive of the response vector with respect to stimulus coordinate m. Our stimulus dimensions are the coordinates of the input vector Rext and output in a network model is R = PN Rext. It is easy to see that ∂mR is given by the mth column of P, and hence I = PT Σ−1P. (105) Generally, if the input depends linearly on a one-dimensional stimulus s, Rext(s) = Rext,0 + s∆Rext, the information about changes in this direction is ∆RT extI∆Rext. As a measure for the information about stimulus changes in all possible directions, we use the trace of I, Tr(I) =(cid:80) i Iii. Covariances in a recurrent network with external noise are Σ = P(D[R] + Σext)PT , where Σext is the diagonal 2 × 2 matrix describing the variance of external input to the two popu- lations. This results in a recurrent linear Fisher Information matrix I r = (D[R] + Σext)−1. (106) This can be compared to a feed-forward scenario with identical transfer matrix, F = P, where Σ = PΣextPT + D[R], so that I f = PT (PΣextPT + D[R])−1P. (107) References [1] van Vreeswijk C, Sompolinsky H. Chaotic balanced state in a model of cortical circuits. Neural Comput. 1998;10(6):1321–71. [2] Shadlen MN, Newsome WT. The variable discharge of cortical neurons: implications for connectivity, computation, and information coding. J Neurosci. 1998;18(10):3870–96. [3] Litwin-Kumar A, Doiron B. Slow dynamics and high variability in balanced cor- Nat Neurosci. 2012;15(11):1498–505. tical networks with clustered connections. doi:10.1038/nn.3220. [4] Renart A, la Rocha JD, Bartho P. The asynchronous state in cortical circuits. Science. 2010;327(5965):587–590. doi:10.1126/science.1179850. [5] Pernice V, Staude B, Cardanobile S, Rotter S. Correlations doi:10.1371/journal.pcbi.1002059. in Neuronal Networks. How Structure Determines PLoS Comput Biol. 2011;7(5):e1002059. [6] Trousdale J, Hu Y, Shea-Brown E, Josi´c K. Impact of network structure and cel- lular response on spike time correlations. PLoS Comput Biol. 2012;8(3):e1002408. doi:10.1371/journal.pcbi.1002408. 39 [7] Goris RLT, Movshon JA, Simoncelli EP. Partitioning neuronal variability. Nat Neurosci. 2014;17(6):858–65. doi:10.1038/nn.3711. [8] Scholvinck ML, Saleem AB, Benucci A, Harris KD, Carandini M. Cortical State Deter- mines Global Variability and Correlations in Visual Cortex. J Neurosci. 2015;35(1):170– 178. doi:10.1523/JNEUROSCI.4994-13.2015. [9] Lin IC, Okun M, Carandini M, Harris KD. The Nature of Shared Cortical Variability. Neuron. 2015;87(3):644–656. doi:10.1016/j.neuron.2015.06.035. [10] Ecker AS, Denfield GH, Bethge M, Tolias AS. On the Structure of Neuronal Population Activity under Fluctuations in Attentional State. J Neurosci. 2016;36(5):1775–1789. doi:10.1523/JNEUROSCI.2044-15.2016. [11] Doiron B, Litwin-Kumar A, Rosenbaum R, Ocker GK, Josi´c K. The mechanics of state- dependent neural correlations. Nat Neurosci. 2016;19(3):383–393. doi:10.1038/nn.4242. [12] Kohn A, Coen-cagli R, Kanitscheider I, Pouget A. Correlations and Neuronal Population Information. Annu Rev Neurosci. 2016;39:237–256. doi:10.1146/annurev-neuro-070815- 013851. [13] Pillow JW, Shlens J, Paninski L, Sher A, Litke AM, Chichilnisky EJ, et al. Spatio- temporal correlations and visual signalling in a complete neuronal population. Nature. 2008;454(7207):995–9. doi:10.1038/nature07140. [14] Vidne M, Ahmadian Y, Shlens J, Pillow JW, Kulkarni J, Litke AM, et al. Modeling the impact of common noise inputs on the network activity of retinal ganglion cells. J Comput Neurosci. 2012;33(1):97–121. doi:10.1007/s10827-011-0376-2. [15] Ecker AS, Berens P, Cotton RJ, Subramaniyan M, Denfield GH, Cadwell CR, et al. State dependence of noise correlations in macaque primary visual cortex. Neuron. 2014;82(1):235–48. doi:10.1016/j.neuron.2014.02.006. [16] Abbott LF, Dayan P. The effect of correlated variability on the accuracy of a population code. Neural Comput. 1999;11(1):91–101. [17] Wilke SD, Eurich CW. Representational accuracy of stochastic neural populations. Neural Comput. 2002;14(1):155–189. doi:10.1162/089976602753284482. [18] Shamir M, Sompolinsky H. Nonlinear population codes. Neural Comput. 2004;16(6):1105–1136. doi:10.1162/089976604773717559. [19] Averbeck BB. Neural correlations, population coding and computation. Nat Rev Neu- rosci. 2006;7(5):358–366. doi:10.1038/nrn1888. [20] Tkacik G, Prentice JS, Balasubramanian V, Schneidman E. Optimal population coding by noisy spiking neurons. P Natl Acad Sci USA. 2010;107(32):14419–24. doi:10.1073/pnas.1004906107. [21] Ecker AS, Berens P, Tolias AS, Bethge M. in populations of diversely tuned neurons. doi:10.1523/JNEUROSCI.2539-11.2011. The effect of noise correlations J Neurosci. 2011;31(40):14272–83. [22] da Silveira RA, Berry MJ. High-Fidelity Coding with Correlated Neurons. PLoS Comput Biol. 2014;10(11):e1003970. doi:10.1371/journal.pcbi.1003970. [23] Hu Y, Zylberberg J, Shea-Brown E. The sign rule and beyond: boundary effects, flexibility, and noise correlations in neural population codes. PLoS Comput Biol. 2014;10(2):e1003469. doi:10.1371/journal.pcbi.1003469. [24] Franke F, Fiscella M, Sevelev M, Roska B, Hierlemann A, da Silveira RA. Struc- tures of Neural Correlation and How They Favor Coding. Neuron. 2016;89(2):409–422. doi:10.1016/j.neuron.2015.12.037. [25] Zylberberg J, Cafaro J, Turner MH, Shea-Brown E, Rieke F. Direction-Selective Cir- cuits Shape Noise to Ensure a Precise Population Code. Neuron. 2016;89(2):369–383. doi:10.1016/j.neuron.2015.11.019. 40 [26] Zohary E, Shadlen MN, Newsome WT. Correlated neuronal discharge rate and Nature. 1994;370(6485):140–3. its implications for psychophysical performance. doi:10.1038/370140a0. [27] Beck J, Bejjanki VR, Pouget A. Insights from a simple expression for linear fisher information in a recurrently connected population of spiking neurons. Neural Comput. 2011;23(6):1484–1502. doi:10.1162/NECO a 00125. [28] Renart A, van Rossum MCW. Transmission of population-coded information. Neural Comput. 2012;24(2):391–407. doi:10.1162/NECO a 00227. [29] Seri`es P, Latham PE, Pouget A. Tuning curve sharpening for orientation selectivity: coding efficiency and the impact of correlations. Nat Neurosci. 2004;7(10):1129–35. doi:10.1038/nn1321. [30] Moreno-Bote R, Beck J, Kanitscheider I, Pitkow X, Latham P, Pouget A. Information- limiting correlations. Nat Neurosci. 2014;17(10):1410–1417. doi:10.1038/nn.3807. [31] Kanitscheider I, Coen-Cagli R, Pouget A. Origin of information-limiting noise correla- tions. P Natl Acad Sci USA. 2015;112(50):E6973–E6982. doi:10.1073/pnas.1508738112. [32] Bathellier B, Ushakova L, Rumpel S. Discrete neocortical dynamics predict behavioral categorization of sounds. Neuron. 2012;76(2):435–49. doi:10.1016/j.neuron.2012.07.008. [33] Hawkes AG. Point spectra of some mutually exciting point processes. J R Stat Soc B. 1971;33(3):438–443. [34] Brody CD. Correlations without synchrony. Neural Comput. 1999;11(7):1537–1551. [35] Grytskyy D, Tetzlaff T, Diesmann M, Helias M. A unified view on weakly Front Comput Neurosci. 2013;7(October):131. correlated recurrent networks. doi:10.3389/fncom.2013.00131. [36] Okun M, Steinmetz NA, Cossell L, Iacaruso MF, Ko H, Barth´o P, et al. Diverse cou- pling of neurons to populations in sensory cortex. Nature. 2015;521(7553):511–515. doi:10.1038/nature14273. [37] Brunel N. Dynamics of sparsely connected networks of excitatory and inhibitory spiking neurons. J Comput Neurosci. 2000;8(3):183–208. [38] Ponce-Alvarez A, Thiele A, Albright TD, Stoner GR, Deco G. Stimulus-dependent variability and noise correlations in cortical MT neurons. P Natl Acad Sci USA. 2013;110(32):13162–7. doi:10.1073/pnas.1300098110. [39] Ostojic S. Two types of asynchronous activity in networks of excitatory and inhibitory spiking neurons. Nat Neurosci. 2014;17(4):594–600. doi:10.1038/nn.3658. [40] Paninski L. Maximum likelihood estimation of cascade point-process neural encoding models. Network - Comp Neural. 2004;15(4):243–262. doi:10.1088/0954-898X/15/4/002. [41] Macke JH, Buesing L, Cunningham JP, Yu BM, Shenoy KV, Sahani M. Empiri- cal models of spiking in neuronal populations. Adv Neur In. 2011;24:1350--1358. doi:10.1.1.230.7630. [42] Rocha J, Doiron B, Reyes A, Shea-brown E, Josi´c K. Correlation between neural spike trains increases with firing rate. Nature. 2007;448(7155):802–6. doi:10.1038/nature06028. [43] Luczak A, Bartho P, Harris KD. Gating of sensory input by spontaneous cortical activity. J Neurosci. 2013;33(4):1684–95. doi:10.1523/JNEUROSCI.2928-12.2013. [44] Cohen MR, Kohn A. Measuring and interpreting neuronal correlations. Nat Neurosci. 2011;14(7):811–9. doi:10.1038/nn.2842. [45] Rothschild G, Nelken I, Mizrahi A. Functional organization and population dy- namics in the mouse primary auditory cortex. Nat Neurosci. 2010;13(3):353–60. doi:10.1038/nn.2484. 41
1606.02627
1
1606
2016-06-08T16:33:41
Brains on Beats
[ "q-bio.NC" ]
We developed task-optimized deep neural networks (DNNs) that achieved state-of-the-art performance in different evaluation scenarios for automatic music tagging. These DNNs were subsequently used to probe the neural representations of music. Representational similarity analysis revealed the existence of a representational gradient across the superior temporal gyrus (STG). Anterior STG was shown to be more sensitive to low-level stimulus features encoded in shallow DNN layers whereas posterior STG was shown to be more sensitive to high-level stimulus features encoded in deep DNN layers.
q-bio.NC
q-bio
Brains on Beats Umut Gu¸clu1, Jordy Thielen1, Michael Hanke2,3, and Marcel A. J. van Gerven1 1Radboud University, Donders Institute for Brain, Cognition and Behaviour, Nijmegen, the Netherlands 2Psychoinformatics lab, Institute of Psychology, Otto-von-Guericke University, Magdeburg, Germany 3Center for Behavioral Brain Sciences, Magdeburg, Germany Abstract We developed task-optimized deep neural networks (DNNs) that achieved state-of-the-art performance in different evaluation scenarios for auto- matic music tagging. These DNNs were subsequently used to probe the neural representations of music. Representational similarity analysis re- vealed the existence of a representational gradient across the superior temporal gyrus (STG). Anterior STG was shown to be more sensitive to low-level stimulus features encoded in shallow DNN layers whereas poste- rior STG was shown to be more sensitive to high-level stimulus features encoded in deep DNN layers. 1 Introduction The human sensory system is devoted to the processing of sensory information to create our perception of the environment [1]. Sensory cortices are thought to encode a hierarchy of ever more invariant representations of the environment [2]. An important question thus is what sensory information is processed as one traverses the sensory pathways from the primary sensory areas to higher sensory areas. This question is the subject matter of sensory neuroscience. The majority of the work on auditory cortical representations has been lim- ited to hand-designed low-level representations such as spectro-temporal mod- els [3], spectro-location models [4], timbre, rhythm, tonality [5, 6, 7] and pitch [8] or high-level representations such as music genre [9] and sound categories [10]. For example, Santoro et al. [3] found that a joint frequency-specific modulation transfer function predicted observed fMRI activity best compared to frequency- nonspecific and independent models. They showed specificity to fine spectral modulations along Heschl's gyrus (HG) and anterior superior temporal gyrus (STG), whereas coarse spectral modulations were mostly located posterior- laterally to HG, on the planum temporale (PT), and STG. Preference for slow temporal modulations was found along HG and STG, whereas fast temporal modulations were observed on PT, and posterior and medially adjacent to HG. Also, it has been shown that activity in STG, somatosensory cortex, the default 1 mode network, and cerebellum are sensitive to timbre, while amygdala, hip- pocampus and insula are more sensitive to rhythmic and tonality features [5, 7]. However these efforts have not yet provided a complete algorithmic account of sensory processing in the auditory system. Since their resurgence, deep neural networks (DNNs) coupled with functional magnetic resonance imaging (fMRI) have provided a powerful approach to form and test alternative hypotheses about what sensory information is processed in different brain regions. On one hand, a task-optimized DNN model learns a hi- erarchy of nonlinear transformations in a supervised manner with the objective of solving a particular task. On the other hand, fMRI measures local changes in blood-oxygen-level dependent hemodynamic responses to sensory stimulation. Subsequently, any subset of the DNN representations that emerge from this hier- archy of nonlinear transformations can be used to probe neural representations by comparing DNN and fMRI responses to the same sensory stimuli. Con- sidering that the sensory systems are biological neural networks that routinely perform the same tasks as their artificial counterparts, it is not inconceivable that DNN representations are suitable for probing neural representations. Indeed, this approach has been shown to be extremely successful in visual neuroscience. To date, several task-optimized DNN models were used to accu- rately model visual areas on the dorsal and ventral streams [11, 12, 13, 14, 15, 16, 17, 18], revealing representational gradients where deeper neural network layers map to more downstream areas along the visual pathways [19, 20]. Recently, [21] has shown that deep neural networks trained to map speech excerpts to word labels could be used to predict brain responses to natural sounds. Here, deeper neural network layers were shown to map to auditory brain regions that were more distant from primary auditory cortex. In the present work we expand on this line of research where our aim is to model how the human brain responds to music. We achieve this by probing neu- ral representations of music features across the superior temporal gyrus using a deep neural network optimized for music tag prediction. We use the represen- tations that emerged after training a DNN to predict tags of musical excerpts as candidate representations for different areas of STG in representational sim- ilarity analysis. We show that different DNN layers correspond to different locations along STG such that anterior STG is shown to be more sensitive to low-level stimulus features encoded in shallow DNN layers whereas posterior STG is shown to be more sensitive to high-level stimulus features encoded in deep DNN layers. 2 Materials and Methods 2.1 MagnaTagATune Dataset We used the MagnaTagATune dataset [22] for DNN estimation. The dataset contains 25.863 music clips. Each clip is a 29-seconds-long excerpt belonging to one of the 5223 songs, 445 albums and 230 artists. The clips span a broad range of genres like Classical, New Age, Electronica, Rock, Pop, World, Jazz, Blues, Metal, Punk, and more. Each audio clip is supplied with a vector of binary annotations of 188 tags. These annotations are obtained by humans playing the two-player online TagATune game. In this game, the two players are either 2 presented with the same or a different audio clip. Subsequently, they are asked to come up with tags for their specific audio clip. Afterward, players view each other's tags and are asked to decide whether they were presented the same audio clip. Tags are only assigned when more than two players agreed. The annotations include tags like 'singer', 'no singer', 'violin', 'drums', 'classical', 'jazz', et cetera. We restricted our analysis on this dataset to the top 50 most popular tags to ensure that there is enough training data for each tag. Parts 1-12 were used for training, part 13 was used for validation and parts 14-16 were used for test. 2.2 Studyforrest Dataset We used the extended studyforrest dataset [23] for representational similarity analysis. The dataset contains fMRI data on the perception of musical genres. Twenty participants (age 21-38 years, mean age 26.6 years), with normal hearing and no known history of neurological disorders, listened to 25 six-second-long, 44.1 kHz music clips. The stimulus set comprised five clips per each of the five following genres: Ambient, Roots Country, Heavy Metal, 50s Rock 'n Roll, and Symphonic. Stimuli were selected according to [9]. The Ambient and Symphonic genres can be considered as non-vocal and the others as vocal. Participants completed eight runs, each with all 25 clips. Ultra-high-field (7 Tesla) fMRI images were collected using a Siemens MAG- NETOM scanner, T2*-weighted echo-planar images (gradient-echo, repetition time (TR) = 2000 ms, echo time (TE) = 22 ms, 0.78 ms echo spacing, 1488 Hz/Px bandwidth, generalized auto-calibrating partially parallel acquisition (GRAPPA), acceleration factor 3, 24 Hz/Px bandwidth in phase encoding di- rection), and a 32 channel brain receiver coil. Thirty-six axial slices were ac- quired (thickness = 1.4 mm, 1.4 × 1.4 mm in-plane resolution, 224 mm field-of- view (FOV) centered on the approximate location of Heschl's gyrus, anterior- to-posterior phase encoding direction, 10% inter-slice gap). Along with the functional data, cardiac and respiratory traces, and a structural MRI were col- lected. In our analyses, we only used the data from the 12 subjects (subjects 1, 3, 4, 6, 7, 9, 12, 14, 15, 16, 17, 18) with no known data anomalies as reported in [23]. The anatomical and functional scans were preprocessed as follows: Func- tional scans were realigned to the first scan of the first run and next to the mean scan. Anatomical scans were coregistered to the mean functional scan. Realigned functional scans were slice-time corrected to correct for the differ- ences in image acquisition times between the slices. Realigned and slice-time corrected functional scans were normalized to MNI space. Finally, a general linear model was used to remove noise regressors derived from voxels unrelated to the experimental paradigm and estimate BOLD response amplitudes [24]. We restricted our analyses to the superior temporal gyrus (STG). 2.3 Deep Neural Networks We developed three task-optimized DNN models for tag prediction. Two of the models comprised five convolutional layers followed by three fully-connected layers (DNN-T model and DNN-F model). The inputs to the models were 96000-dimensional time (DNN-T model) and frequency (DNN-F model) domain 3 representations of six second-long audio signals, respectively. One of the models comprised two streams of five convolutional layers followed by three fully con- nected layers (DNN-TF model). The inputs to the streams were given by the time and frequency representations. The outputs of the convolutional streams were merged and fed into first fully-connected layer. The layers were similar to the AlexNet layers [25] except for the following modifications: • The number of convolutional kernels were halved. • The (convolutional and pooling) kernels and strides were flattened. That is, an n × n kernel was changed to an n2 × 1 kernel and an m × m stride was changed to an m2 × 1 stride. • Local response normalization was replaced with batch normalization [26]. • Rectified linear units were replaced with parametric softplus units with initial α = 0.2 and initial β = 0.5 [27]. • Softmax units were replaced with sigmoid units. We used Adam with parameters α = 0.0002, β1 = 0.5, β2 = 0.999,  = 1e−8 and a mini batch size of 36 [28] to train the models by minimizing the binary cross-entropy loss function. Initial model parameters were drawn from a uniform distribution as described in [29]. Songs in each training mini-batch were randomly cropped to six seconds (96000 samples). The epoch in which the validation performance was the highest was taken as the final model (53, 12 and 12 for T, F and TF models, respectively). The DNN models were implemented in Keras [30]. Once trained, we first tested the tag prediction performance of the models and identified the model with the highest performance. To predict the tags of a 29-second-long song excerpt in the test split of the MagnaTagaTune dataset, we first predicted the tags of 24 six-second-long overlapping segments separated by a second and averaged the predictions. We then used the model with the highest performance for nonlinearly trans- forming the stimuli to eight layers of hierarchical representations for subsequent representational similarity analysis [31]. Note that the artificial neurons in the convolutional layers locally filtered their inputs (1D convolution), nonlinearly transformed them and returned temporal representations per stimulus. These representations were further processed by averaging them over time. In contrast, the artificial neurons in the fully-connected layers globally filtered their inputs (dot product), non-linearly transformed them and returned scalar representa- tions per stimulus. These representations were not further processed. These transformations resulted in n matrices of size m × pi where n is the number of layers (8), m is the number of stimuli (25) and p is the number of artificial neurons in the ith layer (48 or 96, 128 or 256, 192 or 384, 192 or 384, 128 or 256, 4096, 4096 and 50 for i = 1, . . . , 8, respectively). 2.4 Representational Similarity Analysis We used Representational Similarity Analysis (RSA) [31] to investigate how well the representational structures of DNN model layers match with that of the re- sponse patterns in STG. In RSA, models and brain regions are characterized 4 by n × n representational dissimilarity matrices (RDMs), whose elements rep- resent the dissimilarity between the neural or model representations of a pair of stimuli. In turn, computing the overlap between the model and neural RDMs provides evidence about how well a particular model explains the response pat- terns in a particular brain region. Specifically, we performed a region of interest analysis as well as a searchlight analysis by first constructing the RDMs of STG (target RDM) and the model layers (candidate RDM). In the ROI analysis, this resulted in one target RDM per subject and eight candidate RDMs. For each subject, we correlated the upper triangular parts of the target RDM with the candidate RDMs (Spearman correlation). We quantified the similarity of STG representations with the model representations as the mean correlation. For the searchlight analysis, this resulted in 27277 target RDMs (each derived from a spherical neighborhood of 100 voxels) and 8 candidate RDMs. For each subject and target RDM, we correlated the upper triangular parts of the target RDM with the candidate RDMs (Spearman correlation). Then, the layers which re- sulted in the highest correlation were assigned to the voxels at the center of the corresponding neighborhoods. Finally, the layer assignments were averaged over the subjects and the result was taken as the final layer assignment of the voxels. 2.5 Control Models To evaluate the importance of task optimization for modeling STG representa- tions, we compared the representational similarities of the entire STG region and the task-optimized DNN-TF model layers with the representational similarities of the entire STG region and two sets of control models. The first set of control models transformed the stimuli to the following 48- dimensional model representations1: • Mel-frequency spectrum (mfs) representing a mel-scaled short-term power spectrum inspired by human auditory perception where frequencies or- ganized by equidistant pitch locations. These representations were com- puted by applying (i) a short-time Fourier transform and (ii) a mel-scaled frequency-domain filterbank. • Mel-frequency cepstral coefficients (mfccs) representing both broad-spectrum information (timbre) and fine-scale spectral structure (pitch). These rep- resentations were computed by (i) mapping the mfs to a decibel amplitude scale and (ii) multiplying them by the discrete cosine transform matrix. • Low-quefrency mel-frequency spectrum (lq mfs) representing timbre. These representations were computed by (i) zeroing the high-quefrency mfccs, (ii) multiplying them by the inverse of discrete cosine transform matrix and (iii) mapping them back from the decibel amplitude scale. • High-quefrency mel-frequency spectrum (hq mfs) representing pitch. These representations were computed by (i) zeroing the low-quefrency mfccs, (ii) multiplying them by the inverse of discrete cosine transform matrix and (iii) mapping them back from the decibel amplitude scale. 1These are provided as part of the extended studyforrest dataset [23]. 5 The second set of control models were 10 random DNN models with the same architecture as the DNN-TF model, but with parameters drawn from a zero mean and unit variance multivariate Gaussian distribution. 3 Results In the first set of experiments, we analyzed the task-optimized DNN models. The tag prediction performance of the models for the individual tags was defined as the area under the receiver operator characteristics (ROC) curve (AUC). We first compared the mean performance of the models over all tags (Fig- ure 1). The performance of all models was significantly above chance level (p < 0.05, Z-test). The highest performance was achieved by the DNN-TF model (0.8939), followed by the DNN-F model (0.8905) and the DNN-T model (0.8852). To the best of our knowledge, this is the highest tag prediction performance of an end-to-end model evaluated on the same split of the same dataset [32]. The performance was further improved by averaging the predic- tions of the DNN-T and DNN-F models (0.8982) as well as those of the DNN-T, DNN-F and DNN-TF models (0.9007). To the best of our knowledge, this is the highest tag prediction performance of any model (ensemble) evaluated on the same split of the same dataset [33, 32, 34]. For the remainder of the analyses, we considered only the DNN-TF model since it achieved the highest single-model performance. Figure 1: Tag prediction performance of the task-optimized DNN mod- els. Bars show AUCs over all tags for the corresponding task-optimized DNN models. Error bars show ± SE. We then compared the performance of the DNN-TF model for the individ- ual tags (Figure 2). Visual inspection did not reveal a prominent pattern in the performance distribution over tags. The performance was not significantly correlated with tag category or popularity (p > 0.05, Student's t-test). The only exception was that the performance for the positive tags were significantly higher than that for the negative tags (p < 0.05, Z-test). In the second set of experiments, we analyzed how closely the represen- tational geometry of STG is related to the representational geometries of the task-optimized DNN-TF model layers. First, we constructed the candidate RDMs of the layers (Figure 3). Vi- sual inspection revealed similarity structure patterns that became increasingly 6 Figure 2: Tag prediction performance of the task-optimized DNN-TF model. Bars show AUCs for the corresponding tags. Red band shows the mean ± SEM over all tags. prominent with increasing layer depth. The most prominent pattern was the non-vocal and vocal subdivision. Figure 3: RDMs of the task-optimized DNN-TF model layers. Matrix elements show the dissimilarity (1 - Spearman's r) between the model layer representations of the corresponding trials. Matrix rows and columns are sorted according to the genres of the corresponding trials. Second, we performed a region of interest analysis by comparing the reference RDM of the entire STG region with the candidate RDMs (Figure 4). While none of the correlations between the reference RDM and the candidate RDMs reached the noise ceiling (expected correlation between the reference RDM and the RDM of the true model given the noise in the analyzed data [31]), they were all significantly above chance level (p < 0.05, signed-rank test with subject RFX, FDR correction). The highest correlation was found for Layer 1 (0.6811), whereas the lowest correlation was found for Layer 8 (0.4429). Third, we performed a searchlight analysis [35] by comparing the reference RDMs of multiple STG voxel neighborhoods with the candidate RDMs (Fig- ure 5). Each neighborhood center was assigned a layer such that the corre- sponding target RDM and the candidate RDM were maximally correlated. This analysis revealed a systematic change in the mean layer assignments over sub- jects along the STG. They increased from anterior STG to posterior STG such that most voxels in the region of the transverse temporal gyrus were assigned to the shallower layers and most voxels in the region of the angular gyrus were assigned to the deeper layers. The corresponding mean correlations between the target and the candidate RDMs decreased from anterior to posterior STG. 7 Figure 4: Representational similarities of the entire STG region and the task-optimized DNN-TF model layers. Bars show the mean similarity (Spearman's r) of the target RDM and the corresponding candidate RDMs over all subjects. Error bars show ± SE. Red band shows the expected representa- tional similarity of the STG and the true model given the noise in the analyzed data (noise ceiling). Figure 5: Representational similarities of the spherical STG voxel clus- ters and the task-optimized DNN-TF model layers. (A) Mean represen- tational similarities over subjects. (B) Mean layers assignments over subjects. These results show that the representations of increasingly posterior STG voxel neighborhoods can be modeled with the representations of increasingly deeper layers of DNNs optimized for music tag prediction. This observation is in line with the visual neuroscience literature where it was shown that the increasingly deeper layers of DNNs optimized for visual object and action recog- nition can be used to model increasingly downstream visual areas on the ventral and dorsal streams [19, 20]. It also agrees with previous work which used a DNN trained for speech-to-word mapping to show a representational gradient in au- ditory cortex [21]. It would be of particular interest to compare the respective gradients and use the music and speech DNNs as each other's control model such as to disentangle speech- and music-specific representations in auditory cortex. In the last set of experiments, we analyzed the control models. We first constructed the RDMs of the control models (Figure 6). Visual inspection revealed considerable differences between the RDMs of the task-optimized DNN- TF model and those of the control models. We then compared the similarities of the task-optimized candidate RDMs and the target RDM versus the similarities of the control RDMs and the target 8 Figure 6: RDMs of the random DNN model layers (top row) and the baseline models (bottom row). Matrix elements show the dissimilarity (1 - Spearman's r) between the model layer representations of the correspond- ing trials. Matrix rows and columns are sorted according to the genres of the corresponding trials. RDM (Figure 7). The layers of the task-optimized DNN model significantly outperformed the corresponding layers of the random DNN model (∆r = 0.21, p < 0.05, signed-rank test with subject RFX, FDR correction) and the four baseline models (∆r = 0.42 for mfs, ∆r = 0.21 for mfcc, ∆r = 0.44 for lq mfs and ∆r = 0.34 for hq mfs, signed-rank test with subject RFX, FDR correction). Figure 7: Representational similarities of the entire STG region and the task-optimized DNN-TF model versus the representational sim- ilarities of the entire STG region and the control models. Different colors show different control models: Random DNN model, mfs model, mfcc model, lq mfs model and hq mfs model. Bars show mean similarity differences over subjects. Error bars show ± SE. These results show the importance of task optimization for modeling STG representations. This observation also is line with visual neuroscience literature where similar analyses showed the importance of task optimization for modeling ventral stream representations [19, 17]. 9 4 Conclusion We showed that task-optimized DNNs that use time and/or frequency domain representations of music achieved state-of-the-art performance in various eval- uation scenarios for automatic music tagging. Comparison of DNN and STG representations revealed a representational gradient in STG with anterior STG being more sensitive to low-level stimulus features (shallow DNN layers) and posterior STG being more sensitive to high-level stimulus features (deep DNN layers). These results, in conjunction with previous results on the visual and auditory cortical representations, suggest the existence of multiple representa- tional gradients that process increasingly complex conceptual information as we traverse sensory pathways of the human brain. References [1] B. L. Schwartz and J. H. Krantz, Sensation and Perception. SAGE Publi- cations, 2015. [2] J. M. Fuster, Cortex and Mind: Unifying Cognition. Oxford University Press, 2003. [3] R. Santoro, M. Moerel, F. D. Martino, R. Goebel, K. Ugurbil, E. Yacoub, and E. Formisano, "Encoding of natural sounds at multiple spectral and temporal resolutions in the human auditory cortex," PLOS Computational Biology, vol. 10, p. e1003412, jan 2014. [4] M. Moerel, F. D. Martino, K. Ugurbil, E. Yacoub, and E. Formisano, "Processing of frequency and location in human subcortical auditory struc- tures," Scientific Reports, vol. 5, p. 17048, nov 2015. [5] V. Alluri, P. Toiviainen, I. P. Jaaskelainen, E. Glerean, M. Sams, and E. Brattico, "Large-scale brain networks emerge from dynamic processing of musical timbre, key and rhythm," NeuroImage, vol. 59, pp. 3677–3689, feb 2012. [6] V. Alluri, P. Toiviainen, T. E. Lund, M. Wallentin, P. Vuust, A. K. Nandi, T. Ristaniemi, and E. Brattico, "From vivaldi to beatles and back: Predict- ing lateralized brain responses to music," NeuroImage, vol. 83, pp. 627–636, dec 2013. [7] P. Toiviainen, V. Alluri, E. Brattico, M. Wallentin, and P. Vuust, "Cap- turing the musical brain with lasso: Dynamic decoding of musical features from fMRI data," NeuroImage, vol. 88, pp. 170–180, mar 2014. [8] R. D. Patterson, S. Uppenkamp, I. S. Johnsrude, and T. D. Griffiths, "The processing of temporal pitch and melody information in auditory cortex," Neuron, vol. 36, pp. 767–776, nov 2002. [9] M. Casey, J. Thompson, O. Kang, R. Raizada, and T. Wheatley, "Popula- tion codes representing musical timbre for high-level fMRI categorization of music genres," in MLINI, 2011. 10 [10] N. Staeren, H. Renvall, F. D. Martino, R. Goebel, and E. Formisano, "Sound categories are represented as distributed patterns in the human auditory cortex," Current Biology, vol. 19, pp. 498–502, mar 2009. [11] D. L. K. Yamins, H. Hong, C. F. Cadieu, E. A. Solomon, D. Seibert, and J. J. DiCarlo, "Performance-optimized hierarchical models predict neural responses in higher visual cortex," Proceedings of the National Academy of Sciences, vol. 111, pp. 8619–8624, may 2014. [12] P. Agrawal, D. Stansbury, J. Malik, and J. L. Gallant, "Pixels to vox- els: Modeling visual representation in the human brain," arXiv:1407.5104, 2014. [13] S.-M. Khaligh-Razavi and N. Kriegeskorte, "Deep supervised, but not un- supervised, models may explain IT cortical representation," PLOS Com- putational Biology, vol. 10, p. e1003915, nov 2014. [14] C. F. Cadieu, H. Hong, D. L. K. Yamins, N. Pinto, D. Ardila, E. A. Solomon, N. J. Majaj, and J. J. DiCarlo, "Deep neural networks rival the representation of primate IT cortex for core visual object recognition," PLOS Computational Biology, vol. 10, p. e1003963, dec 2014. [15] T. Horikawa and Y. Kamitani, "Generic decoding of seen and imagined objects using hierarchical visual features," arXiv:1510.06479, 2015. [16] R. M. Cichy, A. Khosla, D. Pantazis, A. Torralba, and A. Oliva, "Deep neural networks predict hierarchical spatio-temporal cortical dynamics of human visual object recognition," arXiv:1601.02970, 2016. [17] D. Seibert, D. L. Yamins, D. Ardila, H. Hong, J. J. DiCarlo, and J. L. Gardner, "A performance-optimized model of neural responses across the ventral visual stream," bioRxiv, 2016. [18] R. M. Cichy, A. Khosla, D. Pantazis, and A. Oliva, "Dynamics of scene representations in the human brain revealed by magnetoencephalography and deep neural networks," NeuroImage, apr 2016. [19] U. Gu¸clu and M. A. J. van Gerven, "Deep neural networks reveal a gradi- ent in the complexity of neural representations across the ventral stream," Journal of Neuroscience, vol. 35, pp. 10005–10014, jul 2015. [20] U. Gu¸clu and M. A. J. van Gerven, "Increasingly complex representations of natural movies across the dorsal stream are shared between subjects," NeuroImage, dec 2015. [21] A. Kell, D. Yamins, S. Norman-Haignere, and J. McDermott, "Speech- trained neural networks behave like human listeners and reveal a hierarchy in auditory cortex," in COSYNE, 2016. [22] E. Law, K. West, M. Mandel, M. Bay, and J. S. Downie, "Evaluation of algorithms using games: The case of music tagging," in ISMIR, 2009. 11 [23] M. Hanke, R. Dinga, C. Hausler, J. S. Guntupalli, M. Casey, F. R. Kaule, and J. Stadler, "High-resolution 7-tesla fMRI data on the perception of musical genres – an extension to the studyforrest dataset," F1000Research, jun 2015. [24] K. N. Kay, A. Rokem, J. Winawer, R. F. Dougherty, and B. A. Wandell, "GLMdenoise: a fast, automated technique for denoising task-based fMRI data," Frontiers in Neuroscience, vol. 7, 2013. [25] A. Krizhevsky, I. Sutskever, and G. E. Hinton, "ImageNet classification with deep convolutional neural networks," in NIPS, 2012. [26] S. Ioffe and C. Szegedy, "Batch normalization: Accelerating deep network training by reducing internal covariate shift," arXiv:1502.03167, 2015. [27] J. M. McFarland, Y. Cui, and D. A. Butts, "Inferring nonlinear neuronal computation based on physiologically plausible inputs," PLOS Computa- tional Biology, vol. 9, p. e1003143, jul 2013. [28] D. Kingma and J. Ba, "Adam: A method for stochastic optimization," arXiv:1412.6980, 2014. [29] X. Glorot and Y. Bengio, "Understanding the difficulty of training deep feedforward neural networks," in AISTATS, 2010. [30] F. Chollet, "Keras." https://github.com/fchollet/keras, 2015. [31] N. Kriegeskorte, "Representational similarity analysis – connecting the branches of systems neuroscience," Frontiers in Systems Neuroscience, 2008. [32] S. Dieleman and B. Schrauwen, "End-to-end learning for music audio," in ICASSP, 2014. [33] S. Dieleman and B. Schrauwen, "Multiscale approaches to music audio feature learning," in ISMIR, 2013. [34] A. van den Oord, S. Dieleman, and B. Schrauwen, "Transfer learning by supervised pre-training for audio-based music classification," in ISMIR, 2014. [35] N. Kriegeskorte, R. Goebel, and P. Bandettini, "Information-based func- tional brain mapping," Proceedings of the National Academy of Sciences, vol. 103, pp. 3863–3868, feb 2006. 12
1601.01580
1
1601
2016-01-07T16:02:05
Topological analysis of the connectome of digital reconstructions of neural microcircuits
[ "q-bio.NC", "math.AT" ]
A recent publication provides the network graph for a neocortical microcircuit comprising 8 million connections between 31,000 neurons (H. Markram, et al., Reconstruction and simulation of neocortical microcircuitry, Cell, 163 (2015) no. 2, 456-492). Since traditional graph-theoretical methods may not be sufficient to understand the immense complexity of such a biological network, we explored whether methods from algebraic topology could provide a new perspective on its structural and functional organization. Structural topological analysis revealed that directed graphs representing connectivity among neurons in the microcircuit deviated significantly from different varieties of randomized graph. In particular, the directed graphs contained in the order of $10^7$ simplices {\DH} groups of neurons with all-to-all directed connectivity. Some of these simplices contained up to 8 neurons, making them the most extreme neuronal clustering motif ever reported. Functional topological analysis of simulated neuronal activity in the microcircuit revealed novel spatio-temporal metrics that provide an effective classification of functional responses to qualitatively different stimuli. This study represents the first algebraic topological analysis of structural connectomics and connectomics-based spatio-temporal activity in a biologically realistic neural microcircuit. The methods used in the study show promise for more general applications in network science.
q-bio.NC
q-bio
TOPOLOGICAL ANALYSIS OF THE CONNECTOME OF DIGITAL RECONSTRUCTIONS OF NEURAL MICROCIRCUITS PAWE(cid:32)L D(cid:32)LOTKO∗,1, KATHRYN HESS∗, RAN LEVI∗, MAX NOLTE∗, MICHAEL REIMANN, MARTINA SCOLAMIERO, KATHARINE TURNER, EILIF MULLER, AND HENRY MARKRAM Abstract. A recent publication provides the network graph for a neocor- tical microcircuit comprising 8 million connections between 31,000 neurons [7]. Since traditional graph-theoretical methods may not be sufficient to un- derstand the immense complexity of such a biological network, we explored whether methods from algebraic topology could provide a new perspective on its structural and functional organization. Structural topological analy- sis revealed that directed graphs representing connectivity among neurons in the microcircuit deviated significantly from different varieties of randomized graph. In particular, the directed graphs contained in the order of 107 sim- plices groups of neurons with all-to-all directed connectivity. Some of these simplices contained up to 8 neurons, making them the most extreme neuronal clustering motif ever reported. Functional topological analysis of simulated neuronal activity in the microcircuit revealed novel spatio-temporal metrics that provide an effective classification of functional responses to qualitatively different stimuli. This study represents the first algebraic topological analysis of structural connectomics and connectomics-based spatio-temporal activity in a biologically realistic neural microcircuit. The methods used in the study show promise for more general applications in network science. The Blue Brain Project (BBP) has recently generated the first draft digital reconstruction and simulation of a microcircuit of neurons in the neocortex of a two-week-old rat (Figure 1A) [7]. This reconstruction is made available through the Neocortical Microcircuit Portal (https://bbpnmc.epfl.ch) [11]. Based on sparse anatomical and physiological data for neurons and synapses and on a variety of bi- ologically motivated organizing principles, the complete connectivity between neu- rons belonging to a neocortical microcircuit was digitally reconstructed – a micro- connectome. The structural properties of the reconstruction have been extensively validated against independent data, and simulations of the reconstruction repro- duced multiple in vitro and in vivo experiments without adjusting any parameter, further validating its biological accuracy. In this article we apply methods from topology to the analysis of 42 variants of the digital reconstruction, grouped in six sets of seven microciruits each. The first five sets of microcircuits take into account biological variability in layer heights, proportions of cell types, and cell densities from five individual rats, while the sixth set is based on the average reconstruction across the five individuals. To form each neocortical microcircuit. Key words and phrases. Topology, directed flag complex, Betti number, Euler characteristic, ∗co-first author and corresponding author. (1) Partial support provided by the Advanced Grant of the European Research Council GUDHI (Geometric Understanding in Higher Dimensions). 1 2 D(cid:32)LOTKO ET AL. set of microcircuits, seven statistically varying instantiations of the microcircuit were reconstructed [12]. The 42 microcircuits are therefore all distinct, though the degree of resemblance within each set is higher than that between sets. The structural connectivity of each reconstructed microcircuit can be represented as a directed graph with approximately 3 × 104 vertices and 8 × 106 edges, while its functional connectivity can be represented as a time series of subgraphs formed by functionally effective connections. Our topological analysis of the detailed structural and functional connectivity of these 42 neural microcircuits led to a number of surprising observations. Firstly, we found that the distribution of directed cliques (directed all-to-all connected subsets) of neurons by size is highly significantly different from both that in Erdos-R´enyi random graphs with the same number of vertices and the same average connection probability and that in more sophisticated random graphs, constructed either by taking into account distance-dependent probabilities varying within and between cortical layers or morphological types of neurons, or according to Peters’ Rule [9], [10] (Figure 1D). In particular, we found that directed cliques of up to eight neurons are highly prominent motifs in the reconstructed microcircuits: the average microcircuit incorporates approximately 108 3-cliques and 4-cliques, approximately 107 5-cliques, approximately 105 6-cliques, and approximately 103 7-cliques. Taking the alternating sum of the numbers of directed cliques of various sizes, we computed the Euler characteristic (EC) [5] of the 42 reconstructed microcircuits, obtaining in each case a value on the order of 107, indicating a preponderence of directed cliques consisting of an odd number of neurons (Figure 2). Another topological metric that we considered in this analysis are the Betti numbers (SI, Supplementary Text, ST1.3) associated to a graph via its directed flag complex (Figure 1C). These are a sequence of natural numbers β0, β1, β2, ... that measure the higher-order organizational complexity of the network, detecting “cyclic” chains of intersecting directed cliques. For each graph considered here we determined its homological dimension, i.e., the maximum n such that βn (cid:54)= 0. We showed that the reconstructed microcircuits have homological dimension 5 (Figure 2D), whereas the random graphs considered have homological dimension at most 4, strongly indicating that the microcircuits possess a higher degree of organizational complexity than the random graphs. Topological methods also enabled us to distinguish functional responses to differ- ent input patterns fed into the microcircuit through thalamo-cortical connections. We ran simulations of neural activity in one of the reconstructed microcircuits dur- ing one second, over the course of which a given stimulus was applied every 50 ms (Figure 3). We then binned the output of the simulations by 5 ms timesteps and associated to each timestep a transmission-response graph, the vertices of which are all of the neurons in the microcircuit and the edges of which encode connections in the microcircuit whose activity in that time step leads to firing of the postsynaptic neuron (Figure 4). The size of the time bins and the precise rule for formation of the transmission-response graph for each time bin are biologically motivated, as explained in more detail in the Supplementary Methods section (SI, Supplementary Methods, SM1). From the time series of transmission-response graphs for each of 20 trials of two different stimuli (called Circle and Point for geometric reasons (Figure 4A), we de- rived time series of two non-topological metrics (mean firing rate and number of TOPOLOGICAL ANALYSIS OF NEURAL MICROCIRCUITS 3 Figure 1. (A) A sparse visualization of the microcircuit (soma and dendrites only). Morphological types are color-coded, with m-types in the same layer having similar colors. (B) Examples of simplices in dimensions 0 through 3. (C) An example of a directed graph and its associated flag complex, in which there is one n- simplex for every directed (n + 1)-clique in the graph. (D) A graph depicting the average number of simplices in each dimension for the flag complexes associated to the reconstructed microcircuit (N- complexes) and the four types of random graphs considered, each with the same number of vertices as the reconstructed microcircuit, where shading indicates standard deviation, which was very small for all except the N-complexes. edges in the transmission-response graph) and five topological metrics (the num- ber of 3-cliques, EC, β0, β1, and β2) and applied a Gaussian Bayes classifier (SI, Supplementary Methods, SM2) to determine how successfully each of the metrics classified the 40 trials in the time bins corresponding to the first two stimulations and in the time bins immediately following those stimulations (Figure 5). In each of those crucial time bins, the metrics that were most successful at classification the number of 3-cliques (denoted 2D in the figure), β2, and, in one case, the Euler characteristic (Figure 2A). We expect the methods applied here will prove useful for studying networks in general. ABCDSimplex count0 simplex1 simplex3 simplex2 simplexoriented graphoriented flag complexN-complexesER-complexesP-complexesM-complexesL-complexes 4 D(cid:32)LOTKO ET AL. Figure 2. (A) An oriented simplicial complex consisting of eight 2-simplices glued together along their 1-dimensional faces, together with a table of its Betti numbers and numbers of simplices in dimensions 0,1, and 2 and a computation illustrating that the Euler characteristic can be computed as the alternating sum of the Betti numbers or the simplex counts. (B) Graph depicting the average Euler characteristic of the reconstructed microcircuit (N-complexes) and of each of the types of random graph consid- ered, where the whisker indicates standard deviation, which was very small, except for N-complexes and P-complexes. (C) Box- and-whisker plots depicting the Euler characteristics of 35 recon- structed microcircuits, seven for each individual rat. (D) Box-and- whisker plots depicting the 5th Betti number of 35 reconstructed microcircuits, seven for each individual rat. 1. Structural topology We computed the binary adjacency matrices of all 42 digitally reconstructed microcircuits and then generated the associated directed flag complexes (SI, Sup- plementary Text, ST1.2), which are oriented simplicial complexes encoding the connectivity of all orders of the underlying directed graph: to each directed n- clique (SI, Supplementary Text, ST1.2) in the underlying graph corresponds to an oriented (n − 1)-simplex in the flag complex, and the faces of a simplex correspond ABDCEC=1-0+1=6-12+8=20120-simplices1-simplices2-simplices1016128 TOPOLOGICAL ANALYSIS OF NEURAL MICROCIRCUITS 5 to the directed subcliques of its associated directed clique (Figure 1 B and C). For each neuron in the microcircuit, there is a vertex in the underlying directed graph that is labelled with the unique global identification number (GID) of the neuron. The (j, k)-coefficient of the structural adjacency matrix is 1 if and only if there is a directed connection in the microcircuit from the neuron with GID j to the neuron with GID k. We refer to this adjacency matrix as the structural matrix of the mi- crocircuit and to its associated directed flag complex as a neocortical microcircuit complex or N-complex . Having computed each of the 42 N-complexes, we counted the simplices in each dimension. For comparison with non-biological matrices, we generated five Erdos- R´enyi random graphs [4] of a comparable size (31,000 vertices) and connection probability 0.8%, the same as the average arising from the structural matrices of the microcircuits (SI, Supplementary Methods SM3.1). We refer to the associated directed flag complexes as ER-complexes. To have a more biological control, we also generated 20 adjacency matrices, given by partly randomizing the structural matrix of one of the average microcir- cuits, taking into account its biologically meaningful division into six layers in 10 cases and into 55 morphological neuron types (m-types) [7] in 10 cases. The ran- domization was carried out so that the distance-dependent connection probability for all pairs of layers (respectively, pairs of m-types) was identical to that of the original matrix, i.e., for each pair of layers (respectively, m-types) the number of connections between them was the same as that of the original and for each 25 µm distance bin the number of connections was identical. The matrices are completely random otherwise (SI, Supplementary Methods SM3.2, SM3.3). We call the asso- ciated directed flag complexes L-complexes (respectively, M-complexes). Note that since each m-type is restricted to a fixed layer, the M-complex should retain more of the structure of the original N-complex than the L-complex. Our final and most biological control consisted in the generation of 10 connectivity matrices for 31,000 neurons according to Peters’ Rule [9], [10] for which the associated directed flag complexes are called P-complexes (SI, Supplementary Methods SM3.4). Having carried out the computations for 10 control matrices out of each randomized set of 20, the very small variance in the results convinced us that no further computations should be needed. The resulting distribution of simplices displayed highly consistent behavior among the N-complexes, all of which we computed, with a small variation among the sam- ples arising from different rats. Note that Figure 1 represents the analysis only of the seven N-complexes arising from the average reconstruction because the randomiza- tions are based on those microcircuits. The ER-complexes showed almost identical behavior among the different instances, as did the L-complexes, M-complexes, and P-complexes. On the other hand, the N-complexes exhibited remarkably different distributions from the various random complexes (Figure 1 D), with much greater numbers of simplices and simplices of significantly higher dimension. We computed the Euler characteristic of all N-complexes, as well as that of the various random complexes, obtaining large positive values in all cases, due to the predominance of even-dimensional (particularly 2-dimensional) simplices. The Betti numbers (SI, Supplementary Text, ST1.2) of a simplicial complex pro- vide a much finer and more sophisticated measure of its organizational complexity than the dimension-wise simplex count or the Euler characteristic. The n-th Betti 6 D(cid:32)LOTKO ET AL. Figure 3. (A) Average firing rate (top-down projection) in the stimulated microcircuit, plotted during the first 35 ms after the first stimulation at t=0 ms in the Point vs. Circle experiment. (B) Raster plots of the same 500 neurons randomly picked from layer 4, for two trials of the circle stimulus. (C) Population PSTH of all neurons in the microcircuit for three trials of the Circle stimulus. (D) Mean firing rate of the Circle and Point stimuli, between t and t + 5 ms, where light shading indicates the standard deviation and dark shading the error of the mean. number, βn, counts the number of chains of simplices intersecting along faces to create an “n-dimensional hole” in the complex, which requires a certain degree of organization among the simplices. On the other hand, computation of the Betti numbers is much more expensive than that of the directed flag complex of a di- rected graph or its Euler characteristic. In fact, the sheer size of the complexes we considered here made it practically impossible to do so on a computer with 256 GB of RAM. We succeeded in computing the highest nonzero Betti numbers of the N- complexes, however, by restricting our attention to the 5-th and 6-th coskeleta (SI, Supplementary Text, ST1.2). The top Betti number in all N-complexes appeared in dimension 5, with β5 varying between 1 and 80 (Figure 2D). By contrast, βn = 0 for all n > 3 for all ER-complexes and P-complexes considered, while βn = 0 for all n > 4 for all L-complexes and M-complexes. Moreover β4 varies between 0 and 6 for all L-complexes and M-complexes, so that these Betti numbers are almost negligible. 2. Functional topology We tested our methods on active microcircuits as well. In an experiment that we call the Point vs. Circle test, we activated in a simulation the incoming thalamo- cortical fibers of one of the average that the stimulated fibers formed first a point shape, then a circle shape [7]. The size of the point shape was chosen such that the average firing rate of the neurons was essentially the same as for the circle shape, and in both cases the fibers were activated regularly and synchronously with a frequency of 20 Hz for one second, similar to the whisker deflection approximation in [7, Figure 17A]. We performed 20 trials of each stimulus (Figure 3). The trials of -5-0 ms0-5 ms5-10 ms10-15 ms15-20 ms20-25 ms25-30 ms30-35 ms0 Hz> 60 HzCirclePointA050100150t (ms)051015202530354045Mean FR (Hz)CirclePointDuuuuuStimulationB606060000Population FR (Hz)150100500CCircle trials - example population PSTHsCircle trials - L4 example raster plots15010050015001500Neurons TOPOLOGICAL ANALYSIS OF NEURAL MICROCIRCUITS 7 each stimulus exhibit biological trial-to-trial variability in the neural response, due to the stochasticity of the synapse models and of some of the ion channel models. The aim of this experiment was to determine whether our topological methods were able to classify the two different stimuli, the point and the circle better than the firing rate, which is largely overlapping for the first two stimulations (see Figure 3D). After a systematic analysis to determine the appropriate time bin size and con- ditions for probable spike transmission from one neuron to another (SI, Supplemen- tary Methods, SM1.4), we divided the activity of the microcircuit into 5 ms time bins for 1 second after the initial stimulation and recorded for each 0 ≤ n < 200 a functional connectivity matrix A(n) for the times between 5n ms and 5(n + 1) ms. The (j, k)-coefficient of the binary matrix A(n) is 1 if and only if the following three conditions are satisfied, where sj i denotes the time of the i-th spike of neuron j. (1) The (j, k)-coefficient of the structural matrix is 1, i.e., there is a structural (2) There is some i such that 5n ms ≤ sj connection from the neuron with GID j to the neuron with GID k. i < 5(n + 1) ms, i.e., the neuron with l − sj i < 7.5 ms, i.e., the neuron with GID j spikes in the n-th time bin. (3) There is some l such that 0 ms < sk GID k spikes after the neuron with GID j, within a 7.5 ms interval. We call the matrices A(n) transmission-response matrices, as it is reasonable to assume that the spiking of neuron k is influenced by the spiking of neuron j under conditions (1)–(3) above. The goal of the Point vs. Circle test was to determine whether topological met- rics, such as simplex counts, Betti numbers and Euler characteristic, could classify correctly two groups of stimuli of a similar nature and whether these metrics con- tain more information than the mean firing rate. In Figure 4C we provide plots of the time series of the average zeroth, first, and second Betti numbers, of the aver- age numbers of 1- and 2-simplices, and of the average Euler characteristic for 20 trials of each stimulus. We applied a Gaussian Bayes classifier (SI, Supplementary Methods, SM2) to each metric in each time bin, to determine their success rate at classifying the various trials of the stimuli. To compare, we also classified the stimuli according to the mean firing rates. To allow for a fair comparison, we used three mean firing rates (between t to t + 5, t + 5 to t + 10, and t + 10 to t + 15 ms) for the classification at each time step t, since the transmission-response edges for time step t are based on information from up to t + 12.5 ms. As illustrated by Figure 5 A, none of the metrics considered, topological or otherwise, succeeded very well at classifying the stimuli for times between 10 ms and 50 ms after the initial stimulation, which is not surprising given the strong similarity between the spatial propagation of activity of the two stimuli during this period (Figure 3). On the other hand, in the very first time bin, immediately after the initial stimulation, the 1- and 2-dimensional simplex counts and β1 and β2 all classify very well. In the second time bin the 2-dimensional simplex count and β2 continue to classify very well, and the Euler characteristic classifies even better. Immediately after the second stimulation, from 50 ms to 55 ms after the initial stimulation, none of the metrics performs very well, but the 2-dimensional simplex count and β2 still have the highest success rate. In the next time bin, from 55 ms to 60 ms after the initial stimulation, the 2-dimensional simplex count and β2 8 D(cid:32)LOTKO ET AL. Figure 4. (A) Schematic representation of the transmission- response paradigm: there will be an edge from j to k in the graph associated to particular time bin if and only if there is a physical connection from neuron j to neuron k, neuron j fires in the time bin, and neuron k fires at most 7.5 ms after the firing of neuron j. Here, shading indicates the error of the mean.(B) Schematic representation of those firing patterns involving a presynaptic and a postsynaptic neuron that lead to an edge in the transmission- response graph, with a red block indicating successful transmission and a white block indicating lack of transmission. (C) Time series plots of the average value of the metrics 1D (number of 1-simplices), 2D (number of 2-simplices), β0 (the zeroth Betti number, i.e., the number of connected components), β1 (the first Betti number), β2 (the second Betti number), and EC (the Euler characteristic) for the Circle and Point stimuli. Here, shading indicates the error of the mean. again classify very well and are the only metrics to do so. In all of these cases, the topological metrics far outperform the metric based on firing rate. 3. Discussion We have introduced topological analysis of directed graphs encoding structural or functional connectivity of digital reconstructions of neural microcircuits. We showed in particular that these directed graphs differed significantly from random graphs of both Erdos-R´enyi-type and types taking into account biologically con- strained, distance-dependent connection probabilities. The topological analysis re- vealed not only the existence of high-dimensional simplices representing the most extreme form of circuit “motifs” - all-to-all connectivity within a set of neurons - that have so far been been reported for brain tissue, but also that there are a surpris- ingly huge number of these structures. We established moreover that topological methods effectively distinguish functional responses to distinct thalamic stimuli, introducing a new measure of the spatio-temporal activity responses generated by neural tissue. The results of our topological analysis of biologically realistic digital 1DStructural connectionTransmission, no responseTransmission-responsePrePostStep∆t2 = 7.5 ms∆t1 = 5 ms1234ABCt (ms)01234561e5t (ms)01234567891e5CirclePointt (ms)1.82.02.22.42.62.83.03.21e4t (ms)0.00.20.40.60.81.01.21e5020406080100120140t (ms)0.00.51.01.52.02.53.01e5020406080100120140t (ms)−0.50.00.51.01.52.02.51e52Dβ1ECβ0β2 TOPOLOGICAL ANALYSIS OF NEURAL MICROCIRCUITS 9 Figure 5. (A) Times series plot for the first 80 ms of the 40 trials of the percentage of correct classifications performed by a Gauss- ian Bayes classifier based on each of the metrics FR (sequences of mean firing rates over three consecutive time bins), 1D (number of 1-simplices), 2D (number of 2-simplices), β0 (the zeroth Betti number, i.e., the number of connected components), β1 (the first Betti number), β2 (the second Betti number), and EC (the Eu- ler characteristic). (B) Graphs depicting the percentage of correct classifications performed by a Gaussian Bayes classifier based on each of the metrics in four particularly important time bins: from 0 to 5 ms (immediately after the initial stimulation), from 5 to 10 ms, from 50 to 55 ms (the time bin immediately after the second stimulation), and from 55 to 60 ms. reconstructions provide a convincing argument for considering topology as a useful mathematical tool for analyzing the structural and functional connectome of neural circuits. Our results lead naturally to many new questions, most notably concerning the biological significance of the high-dimensional simplices and homology classes we have discovered in the digitally reconstructed neocortical microcircuits. We intend to explore these questions in future studies. In particular we hypothesize that the time series of different topological metrics could reveal an evolving spatio-temporal code that goes beyond either rate or timing information to one that incorporates the structural organization. Such metrics could yield a deeper understanding of how 70%80%90%100%Correct classi(cid:31)cations0-5 ms70%80%90%100%5-10 ms70%80%90%100%Correct classi(cid:31)cations70%80%90%100%FR1D2Dβ0β1β2ECFR1D2Dβ0β1β2ECFR1D2Dβ0β1β2ECFR1D2Dβ0β1β2EC50-55 ms55-60 msABFR1D2D012ECβββ01020304050607050%60%70%80%90%100% 10 D(cid:32)LOTKO ET AL. the structural organization constrains emergent functional states. Age-dependent changes in such digital reconstructions may help reveal even more complex topo- logical structures with development, and changes introduced by synaptic plasticity may reveal structures associated with learning and memory. We expect the topological approach to studying directed graphs that we im- plement here will also prove useful in applications of network science outside of neuroscience, in the study of networks exhibiting intricate directed connectivity patterns, such as gene and protein networks, VLSI circuits, and electrical grids. The obvious utility of the directed flag complex in these applications may also encourage theorists to establish results analogous to those established by Kahle concerning Betti numbers of undirected flag complexes of random graphs [6]. 4. Materials and methods 4.1. Computation of flag complexes and their Betti numbers. We represent the directed flag complex of a directed graph by a reference-based data structure, using vectors to store the references to the simplices in the simplicial complex. The required storage space grows linearly with the number of vertices and with the number of edges. A publicly available C++ implementation of the code will be available on http://neurotop.gforge.inria.fr/. All homology computations carried out for this paper were made with F2 coefficients, using the boundary matrix re- duced by an algorithm from the PHAT [2] library. For further details, please see (SI, Supplementary Text, ST2). 4.2. The Point vs. Circle experiment. The stimulated reconstructed microcir- cuit is innervated by 310 VPM fibers, whose horizontal centers of innervation are evenly distributed over the microcircuit (one fiber per mini-column). It is therefore possible to activate the microcircuit with topographically different stimuli by se- lecting only a subset of these 310 fibers. Here we used two different stimuli, a point and a circle, which were calibrated by adjusting the respective number of fibers to evoke an overall similar mean firing rate (i.e., close enough to prevent clearly distin- guishing between the two stimuli simply by the mean population firing rate). The microcircuit was stimulated by synchronous spikes, similar to the whisker deflection experiment described by Markram et al. (2015). The point stimulus consisted of synchronous spikes in the 46 neighboring fibers of the center of the microcircuit, whereas the circle stimulus involved 56 fibers near the periphery of the microcircuit. The stimulation was repeated every 50 ms, but only the firing rates after the first two stimulations (at 0 and 50 ms) are overlapping. We used a Gaussian naıve Bayes classifier [8], where we performed 500 classifica- tion trials, randomly choosing 15 trials of each stimulus to be part of the training data, and five trials of each stimulus to be part of the test data. We then obtained the mean ratio of successfully classified test data points using 500 different training and test sets. The classification of the firing rate used the firing rates of three consecutive time bins, to make it a fairer comparison, since the edges may contain firing rate information of more than two time bins, over a range of 12.5 ms. 4.3. Computation of transmission-response matrices. Transmission-response matrices were calculated according to the specifications mentioned above, using a custom-written program in the Python programming language. It combined the TOPOLOGICAL ANALYSIS OF NEURAL MICROCIRCUITS 11 matrix of synaptic connections (structural matrix), constructed as part of the stan- dard reconstruction process of the BBP, with the spiking output of a simulation run and user-defined values for time steps ∆t1 and ∆t2 (5 and 7.5 ms in our analyses). For further details, please see (SI, Supplementary Methods ,SM1). 4.4. Gaussian Bayes classifiers. The Gaussian Bayes classifier minimises the probability of misclassification under the assumption that the distributions are Gaussian. We randomly split the data into training and testing sets. Using the training set we model the distributions of the dot and circle classes by Gaussians N (µdot, σ2 circle) respectively. Assuming a uniform prior and Gaussian distributions, Bayes’ theorem provides a classifier dot) and N (µcircle, σ2 Class(x) = argmax c∈{dot,circle} 1(cid:112)2πσ2 c exp (cid:18)−(x − µc)2 (cid:19) . 2σ2 c 5. Acknowledgments This work was supported by funding from the ETH Domain for the Blue Brain Project (BBP). The BlueBrain IV IBM BlueGene/Q system is financed by ETH Board Funding to the Blue Brain Project and hosted at the Swiss National Super- computing Center (CSCS). We thank Ahmet Bilgili for providing the visualization of the microcircuit in Figure 1. Partial support for P.D. was provided by the GUDHI project, supported by an Advanced Investigator Grant of the European Research Council and hosted by INRIA. M.S. was supported by the NCCR Synapsy of the Swiss National Science Foundation. References [1] U. Bauer, M. Kerber, J. Reininghaus, Clear and Compress: Computing Persistent Homology in Chunks, TopoInVis 2013, in press. [2] U. Bauer, M. Kerber, J. Reininghaus, Phat library, https://code.google.com/p/phat/. [3] T. H. Cormen, C. E. Leiserson, R. L. Rivest and C. Stein, Introduction to Algorithms, The MIT Press, 2001. [4] Erdos, P.; R´enyi, A. On random graphs, I. Publ. Math. Debrecen 6 1959 290–297. [5] A. Hatcher, Algebraic Topology, Cambridge University Press (Available Online), 2002. [6] M. Kahle, Sharp vanishing thresholds for cohomology of random flag complexes, Ann. of Math. (2), 179 (2014), 1085–1107. [7] H. Markram, et al., Reconstruction and simulation of neocortical microcircuitry, Cell, 163 (2015) no. 2, 456-492. [8] Pedregosa, Fabian, et al., Scikit-learn: Machine learning in Python, The Journal of Machine Learning Research 12 (2011): 2825–2830. [9] Peters, A., and Feldman, M.L. The projection of the lateral geniculate nucleus to area 17 of the rat cerebral cortex. I. General description, J. Neurocytol., 5 (1976), 6384. [10] Peters, A., Proskauer, C.C., Feldman, M.L., and Kimerer, L. The projection of the lateral geniculate nucleus to area 17 of the rat cerebral cortex. V. Degenerating axon terminals synapsing with Golgi impregnated neurons, J. Neurocytol., 8 (1979), 331357. [11] S. Ramaswamy, rat et al., The neocortical microcircuit collaboration portal: somatosensory in Neural Circuits, cortex , Frontiers a re- 9 (2015), source http://dx.doi.org/10.3389/fncir.2015.00044. for [12] M. Reimann, J. King, E. Muller, S. Ramaswamy, and H. Markram, An algorithm to predict the connectome of neural microcircuits, Frontiers in Computational Neuroscience (2015) 120, doi:10.3389/fncom.2015.00120. [13] R. Milo, S. Shen-Orr, S. Itzkovitz, N. Kashtan, D. Chklovskii, and U. Alon, Network Motifs: Simple Building Blocks of Complex Networks, Science, 25 October 2002: 298 (5594), 824–827. [DOI:10.1126/science.298.5594.824] 12 D(cid:32)LOTKO ET AL. SI Appendix To accompany “Topological analysis of the connectome of digital reconstructions Pawe(cid:32)l D(cid:32)lotko, Kathryn Hess, Ran Levi, Max Nolte, Michael Reimann, Martina Scolamiero, Katharine Turner, Eilif Muller, Henry Markram of neural microcircuits.” Contents (1) Supplementary Methods (3 pages) (2) Supplementary Text (4 pages) (3) Supplementary Figures (5 pages) Supplementary Methods SM1. Optimization of the parameters for the transmission-response matrices For fixed values of ∆t1 and ∆t2, the corresponding sequence {A(n)}N The transmission-response matrices that allow us to analyze activity in an exper- iment (cf. the section on Functional Topology in the main body of the article) form a sequence depending on two parameters, ∆t1 and ∆t2. The number of matrices in the sequence is the duration of the experiment divided by ∆t1. In other words for a given experiment of duration T and fixed ∆ti, we obtain a sequence of matrices S(∆t1, ∆t2) = {A(n) = A(n, ∆t1, ∆t2)}N n=1, where N is the integer value of T /∆t1. n=1 is ob- tained as follows. The spiking output of the simulation is first converted into lists of spike times, one for each neuron. Standard histogram methods, binning by ∆t1, are applied to each list to determine in which time steps a presynaptic neuron fired. For each time bin in which a particular neuron fired, the exact timing of its first spike in that bin is then compared to the full list of spike times of each neuron it innervates, to ascertain which of them had spiked at most ∆t2 ms after the presy- naptic neuron. (Spiking of a pair of neurons within ∆t2 ms is ignored if they are not structurally connected.) For all pre-postsynaptic pairs satisfying this constraint on spike timing, the corresponding entry in the transmission-response matrix for that time step is set to 1 and all others to 0. More precisely, the (j, k)-coefficient of the binary transmission-response matrix A(n) corresponding to the n-th time bin is 1 if and only if the following three conditions are satisfied, where sj i denotes the time of the i-th spike of neuron j. (1) The (j, k)-coefficient of the structural matrix is 1, i.e., there is a structural connection from the neuron with GID j to the neuron with GID k, so that they form a pre-post synaptic pair. GID k spikes after the neuron with GID j, within a ∆t2 ms interval. Starting with firing data from spontaneous activity in the reconstructed mi- crocircuit, we generated sequences of 20 transmission-response matrices for ∆ti ∈ {1, 2, 5, 10, 20, 50, 100} ms, thus creating 49 such sequences corresponding to every possible choice of the pair (∆t1, ∆t2). We refer to each of these sequences as the true transmission-response matrices corresponding to the pair (∆t1, ∆t2). (2) There is some i such that n∆t1 ms ≤ sj with GID j spikes in the n-th time bin. l − sj (3) There is some l such that 0 ms < sk i < (n + 1)∆t1 ms, i.e., the neuron i < ∆t2 ms, i.e., the neuron with TOPOLOGICAL ANALYSIS OF NEURAL MICROCIRCUITS 13 In the rest of this section, we describe the procedure for optimizing the choice of the time intervals ∆t1 and ∆t2 so that the associated true transmission-response matrices best reflect the actual successful transmission of signals between neurons in the microcircuit. SM1.1. Properties of the transmission-response matrix. The nonzero co- efficients in a transmission-response matrix are a subset of those in the structural matrix. Due to the partly stochastic behavior of the in silico microcircuit, the sub- set will vary even for subsequent applications of the same stimulus. In fact, even an exact repetition of the same conditions will lead to different transmission-response matrices, if the random number generator is seeded differently. It follows that the generation of the transmission-response matrices for a given stimulus should be considered as a stochastic process. With the correct choice of the parameters ∆ti, the matrices should reflect how the microcircuit processes a stimulus and thus take into account parameters of neural processing, such as pre-post synaptic interaction. To find parameters ∆t1 and ∆t2 that maximize the degree to which neural processing is captured by the transmission-response matrices, we first develop a stochastic model for synaptic firing that takes into account neural processing and that depends on ∆t1 and ∆t2. For the purpose of this analysis, we assume that the true transmission-response matrices are compatible with this model. Based upon our model for synaptic firing, we formulate a simplified model that ignores neural processing. For this simplified model and for any choice of parameters ∆t1 and ∆t2, we explain how to obtain transmission-response matrices from actual firing data, by shuffling the firing data appropriately, then applying the algorithm for generating a transmission-response matrix of the previous section. Finally, for each choice of the parameters ∆t1 and ∆t2, we compare the true transmission- response matrices for spontaneous activity in the reconstructed microcircuit to those obtained by the simplified generation process. The parameters that we work with in the main body of the paper are the ∆t1 and ∆t2 that maximize the difference (measured by the ratio of the numbers of ones in the matrices) between the actual transmission-response matrices and those resulting from the simplified model. SM1.2. Stochastic model with neural processing. Fix time intervals ∆t1 and ∆t2. Let A = (aij) denote the structural matrix of a reconstructed microcircuit, and let A(n) = (an ij) denote the transmission-response matrix of the n-th time bin, based on firing data from a trial of simulated activity in the microcircuit, for the given intervals ∆t1 and ∆t2. By Condition (1) above, if an ij = 1 for any n, then aij = 1. It is reasonable to consider A to be static, at least over the time periods considered here. We want to compute the probability that an ij = 1, given that aij = 1, so we need to determine on which parameters and properties this probability depends. According to the definition of transmission-response matrices, a presynaptic and a postsynaptic spike are required for an ij to be 1. To simplify the analysis somewhat, we assume that each neuron ni has a time-dependent, instantaneous firing rate F i(t) that determines spiking probability at time t, i.e., spiking can be described as an inhomogeneous Poisson process. Under this assumption, the expected number mi (t0) of spikes of neuron ni in the interval [t0, t0 + ∆t1] can be computed as ∆t1 (cid:90) t0+∆t1 mi ∆t1 (t0) = t0 F i(u)du. 14 D(cid:32)LOTKO ET AL. ∆t1 ∆t1 ∆t1 K i −mi (t0) denotes the probability that neuron ni spikes at least once in the interval If K i [t0, t0 + ∆t1], then (t0) = 1 − P(cid:0)mi the interval [t0, t0 + ∆t1], then P(cid:0)mi (t0)(cid:1) = 1 − e (t0)(cid:1) is the probability that X = 0.) If the where P(λ) is the Poisson probability mass function with parameter λ at 0. (Recall that if X is a random variable that counts the number of spikes of neuron ni in (t) ≈ F i(t) · ∆t1. Moreover, change in F i(t) is slow compared to ∆t1, then mi 1 − P(λ) ≈ λ for small values of λ. For small enough ∆t1, the expected number (t0) of spikes of neuron ni will certainly be small, and change in F i(t) will be mi slow in compared to ∆t1, so that we may assume that (t0) ≈ F i(t0) · ∆t1. (t0), K i ∆t1 ∆t1 ∆t1 ∆t1 ∆t1 For the postsynaptic spike the situation is more complicated. As there is a causal relation between presynaptic and postsynaptic firing, mediated by synaptic transmission, we need to estimate the conditional probability of at least one post- synaptic spike, given that at least one presynaptic spike occured. Let ni and nj denote neurons such that aij = 1. Let s0 ∈ [t0, t0 + ∆t1] denote the time of the first presynaptic spike in this interval. Let X j (s0) denote the random variable whose value is the number of times neuron nj spiked in the time window [s0, s0 + ∆t2]. Let Y i (t0) denote the random variable whose value is the number of times neuron ni spiked in the time interval [t0, t0 + ∆t1]. We need to estimate the conditional probability ∆t2 ∆t1 P(cid:0)X j (s0) > 0 Y i ∆t1 ∆t2 (t0) > 0(cid:1). The nature of this interaction is very intricate and depends on the identities of the presynaptic and postsynaptic neurons, the spiking history of the presynaptic neuron before s0, and all other synaptic input the postsynaptic neuron received. It can be described as governed by some function Gij modulating the spiking probability of the postsynaptic neuron nj. This function takes as parameters the expected number of spikes of neuron nj in the interval [s0, s0 + ∆t2], the time t0, and the “spiking history” of the presynaptic neuron ni until s0, which we write as a function si∗(t) evaluated at s0, giving rise to the expression P(cid:0)X j (s0) > 0 Y i ∆t1 ∆t2 (t0) > 0(cid:1) = 1 − e −Gij (mj ∆t2 (s0),t0,si∗(s0)). Summarizing the analysis above, the following formula provides a good estimate of the probability that an ij = 1 if aij = 1, for small enough ∆t1 and ∆t2, where s0 denotes the time of the first presynaptic spike in the interval [n∆t1, (n + 1)∆t1] and t0 = n∆t1. (cid:16) ≈ F i(t0) · ∆t1 · Gij(cid:0)F j(s0) · ∆t2, t0, si∗(s0)(cid:1). (t0)(cid:17) ·(cid:16) (s0),t0,si∗(s0))(cid:17) −Gij (mj 1 − e 1 − e P(cid:0)an ij = 1aij = 1(cid:1) = −mi (1) ∆t2 ∆t1 This conditional probability encodes not only the distinctive features of the struc- tural connectivity (via aij) but also the potentially stimulus-dependant neuron- specific firing rates (via F i and F j) and their co-variation. Most crucially, it captures the stimulus-dependent functional modulation of postsynaptic firing by a presynaptic spike as well. We assume that the true transmission-reponse matri- ces capture the actual transmission of spikes according to the model of synaptic firing described by this formula. TOPOLOGICAL ANALYSIS OF NEURAL MICROCIRCUITS 15 SM.1.3. Null hypotheses: no neural processing. We introduce here a sim- plified model of synaptic spiking that is based upon formula [1] but that ignores pre-post synaptic interaction. We then explain how to obtain transmission-response matrices that correspond to this simplified model from firing data arising from sim- ulated activity. We begin by setting each Gij to be the projection onto the first component, ignoring the pre-post synaptic interaction. After this simplification, the approxi- mation obtained in the previous section now reads P (an ij = 1aij = 1) ≈ F i(t0) · F j(s0) · ∆t1 · ∆t2, where s0 denotes the time of the first presynaptic spike in the interval [n∆t1, (n + 1)∆t1] and t0 = n∆t1, as before. Since this drastic simplification neglects the central as- pect of neural computation - pre-post synaptic interaction - it gives rise to control cases for each pair of parameters (∆t1, ∆t2) and each choice of firing rate func- tions F i(t). Comparison of the true transmission-response matrices for each pair of parameters to the corresponding control matrices for the same pair and a specific choice of the functions F i(t) will allow us to determine values for ∆t1 and ∆t2 for which the true transmission-response matrix optimally reflects neural processing. We assume moreover that the individual firing rates consist of a neuron-dependent frequency that is up- or down-regulated by a global time series, i.e., that F i(t) = f (i) · F (t), for some function F (t) and some constant f (i) for each neuron ni. Transmission-response matrices corresponding to this simplified model for fixed ∆t1 and ∆t2, which we call simplified transmission-response matrices, can be generated by first shuffling all recorded spikes from simulated activity in the reconstructed microcircuit, while preserving both the number of spikes per neuron and per time bin, then applying the usual transmission-response matrix generation method. SM.1.4. Optimization of parameters. The difference between the true transmission- response matrices and the control case described above is a consequence of the pre- post synaptic interaction. Comparison with the control case enables us therefore to measure how well that interaction is captured in the true transmission-response matrices. In particular, it is reasonable to optimize the parameters ∆t1 and ∆t2 so that the difference between the true transmission response matrices arising from actual simulation data and those arising in the control cases is maximized, as a maximal difference indicates that the effect of the pre-post synaptic interaction is captured optimally by the true transmission-response matrices. The comparison between the true transmission-response matrices and the control cases was carried out by first producing 20 true transmission-response matrices and 20 simplified transmission-response matrices based on firing data obtained from spontaneous activity in the reconstructed microcircuit for every pair (∆t1, ∆t2), where ∆ti ∈ {1, 2, 5, 10, 20, 50, 100} ms for i = 1, 2. The number of ones in each matrix was then computed and the average taken over each set of 20 matrices. Since no stimulus was applied to the microcircuit, the averages computed are meaningful, since the firing data should be fairly homogeneous across the time bins. The average number of ones in the transmission-response matrix arising from simulated actitivity in the reconstructed microciruit, as a function of ∆t1 and ∆t2, is illustrated in Figure S1. Figure S2 shows the ratio of the average number of ones in the true transmission-response matrices to the average number of ones in the simplified transmission-response matrices, for various values of ∆t1 and ∆t2. 16 D(cid:32)LOTKO ET AL. In all cases we find that the maximum lies between ∆t2 = 5 ms and ∆t2 = 10 ms, leading us to choose to work with ∆t2 = 7.5 ms. For ∆t1 we find a maximum at 50 ms, but we use ∆t1 = 5 ms (for which the maximum ratio is only slightly lower than for ∆t1 = 50 ms) instead to avoid more than one spike per neuron per bin. SM2. Gaussian Bayes classifiers Suppose there is a distribution ρ over R × {c1, c2, . . . , ck}, where {c1, c2, . . . , ck} is a set of class labels. We can project ρ onto each of the coordinates to construct a real-valued random variable X and a class-label-valued random variable Y . We wish to build a classifier C : R → {c1, c2, . . . , ck} which will, for any real number, choose the most likely class to which it might belong. That is, C(x) = argmax c∈{c1,c2,...,ck} P (Y = cX = x), where P (AB) is the probability of A conditional on B and argmaxa∈A f (a) denotes the element a ∈ A such that f (a) is maximal. This element of A will in practice always be unique. Bayes’ theorem states that P (Y = cX = x)P (X = x) = P (X = xY = c)P (Y = c). A Bayesian classifier picks the class with the highest conditional probability, which using Bayes’ theorem is C(x) = argmax c∈{c1,c2,...ck} P (X = xY = c)P (Y = c) . P (X = x) Usually ρ itself is unknown and must be infered from sample data. We then also assume some model distribution to estimate ρ from these samples. The Gaussian Bayes classifier is the Bayes’ classifier under the assumption that the distribution of each separate class is Gaussian. After calculating the means and variances of the sample data within each of the classes separately, we model their respective distributions by the Gaussians N (µci, σ2 ci ). If p(A) denotes the probability density function of A, then P (X = xY = c) P (Y = c)P (X = x) = = p(X = xY = c) P 1(cid:112)2πσ2 c exp (Y = c)p(X = x) (cid:18)−(x − µc)2 (cid:19) P (Y = c) 2σ2 c p(X = x) A common situation, such as in our analysis, is a uniform prior. A uniform prior over {c1, c2, . . . , ck} means P (Y = ci) = 1/k for all i. If we assume a uniform prior, then the factor P (Y =c) p(X=x) is common to all classes and thus does not affect which class achieves the maximum. Thus we get the formula 1(cid:112)2πσ2 c exp (cid:18)−(x − µc)2 (cid:19) . 2σ2 c C(x) = argmax c∈{c1,c2,...ck} TOPOLOGICAL ANALYSIS OF NEURAL MICROCIRCUITS 17 SM3. Randomization of connection matrices and other control cases We created four types of random matrices of sizes and connection probabilities similar to the connectivity matrices of the BBP reconstruction. SM3.1. Generation of Erdos-R´enyi random matrices. For this basic control we first computed the overall connection probability in the reconstruction and found it to be 0.8%. We then generated random, binary square matrices of size 3.1× 104, where 1’s were placed at random off-diagonal in the matrix with probability 0.8%. SM3.2. Randomization preserving the distance-dependent connectivity between layers. Input for this randomization method were the structural matrix and the matrix of pairwise soma distances, both generated as part of the standard BBP reconstruction process. The rows and columns of both matrices were first grouped into N = 6 groups according to the layer of the neuron they correspond to. This effectively partitioned both matrices into N ∗ N = 36 submatrices each. For each pair of submatrices, the soma distances were grouped into bins of size 25µm. Next, in the submatrix corresponding to each distance bin, we first replaced all 1’s by 0’s and then replaced randomly chosen 0’s by 1’s, so that the total number of 1’s was preserved. Creation of autapses, i.e., a connection from a neuron to itself, was avoided by creating a separate bin for distances of 0µm. The result was a connection matrix with the same number of connections between each pair of layers and the same distance-dependent connection probability between pairs of layers, to within 25µm, as the original matrix. SM3.3. Randomization preserving the distance-dependent connectivity between m-types. This randomization method was identical to the preceding randomization, preserving connectivity between layers, except that the neurons were partitioned initially into N = 55 groups of morphological types instead of only six layers. SM3.4. Generation of connection matrices according to Peters’ Rule. For this control case, we started with a connection matrix that placed a connection not just where a synaptic connection was found in the reconstructed microcircuit, but between each pair of neurons whose arbors came within close proximity (closer than 3µm). The resulting matrix had approximately 16 times more connections than the structural matrix. These connections were then pruned randomly with a uniform probability until the same number of connections as in the structural matrix was attained. 18 D(cid:32)LOTKO ET AL. Supplementary Text Supplementary Text To accompany “Topological analysis of the connectome of digital reconstructions Pawe l D(cid:32)lotko, Kathryn Hess, Ran Levi, Max Nolte, Michael Reimann, Martina Scolamiero, Katharine Turner, Eilif Muller, Henry Markram of neural microcircuits.” ST1. The topological toolbox Most of the mathematical methods we describe here are part of the basic toolbox of algebraic topology, though perhaps not as well known in the directed variants presented here. We give a brief account of these concepts for the benefit of the non-expert, and refer to literature for the reader interested in further details. We explain first how to associate to any directed graph a simplicial complex known as its directed flag complex , then recall two types of important invariants of simplicial complexes, which turn out to be very useful for analyzing the digitally reconstructed microcircuits: the Euler characteristic and Betti numbers. We then describe the data structures and algorithms that we implemented in order to con- struct the flag complexes of the directed graphs representing the microcircuits and to compute their Euler characteristics and Betti numbers. ST1.1. Directed graphs. A directed graph G consists of a pair of finite sets (V, E) and a function τ : E → V × V . The elements of the set V are the vertices of G, the elements of E are the edges of G, and the function τ associates with each edge an ordered pair of vertices. The direction of an edge e with τ (e) = (v1, v2) is taken to be from τ1(e) = v1, the source vertex , to τ2(v) = v2, the target vertex . The function τ is required to satisfy the following two conditions. (1) For each e ∈ E, if τ (e) = (v1, v2), then v1 (cid:54)= v2, i.e., there are no loops in the graph. (2) The function τ is injective, i.e., for any pair of vertices (v1, v2), there is at most one edge directed from v1 to v2. A vertex v ∈ G is said to be a sink if τ1(e) (cid:54)= v for all e ∈ E, and a source is if τ2(e) (cid:54)= v for all e ∈ E. To compare two graphs, we require the following notion. A morphism of directed graphs from a directed graph G = (V, E, τ ) to a directed graph G(cid:48) = (V (cid:48), E(cid:48), τ(cid:48)) consists of a pair of set maps α : V → V (cid:48) and β : E → E(cid:48) such that β takes an edge in G with source v1 and target v2 to an edge in G(cid:48) with source α(v1) and target α(v2), i.e., τ(cid:48) ◦ β = (α, α) ◦ τ . Two graphs G and G(cid:48) are isomorphic if there is morphism of graphs (α, β) : G → G(cid:48) such that both α and β are bijections, which we call an isomorphism of directed graphs (Figure S3). A path in a directed graph G consists of a sequence of edges (e1, ..., en) such that for all 1 ≤ k < n, the target of ek is the source of ek+1, i.e., τ2(ek) = τ1(ek+1). The length of the path (e1, ..., en) is n, the number of edges of which the path is composed. If, in addition, target of en is the source of e1, i.e., τ2(en) = τ1(e1), then (e1, ..., en) is an oriented cycle. TOPOLOGICAL ANALYSIS OF NEURAL MICROCIRCUITS 19 ST1.2. Simplicial complexes. An abstract oriented simplicial complex is a col- lection S of finite, ordered sets with the property that if σ ∈ S, then every subset τ of σ is also a member of S. A subcomplex of an abstract oriented simplicial complex is a sub-collection S(cid:48) ⊆ S that is itself an abstract oriented simplicial com- plex. Henceforth, we simplify terminology and usually refer to abstract oriented simplicial complexes merely as simplicial complexes. The elements of a simplicial complex S are called its simplices. A simplicial If σ ∈ S, we complex is said to be finite if it has only finitely many simplices. define the dimension of σ, denoted dim(σ), to be σ − 1, the cardinality of the set σ minus one. If σ is a simplex of dimension n, then we refer to σ as an n-simplex of S. The set of all n-simplices of S is denoted Sn. A simplex τ is said to be a face of σ if τ is a subset of σ of a strictly smaller cardinality. A front face of an n-simplex σ = (v0, ..., vn) is a face τ = (v0, ..., vm) for some m < n. Similarly, a back face of σ is a face τ(cid:48) = (vi, . . . , vn) for some 0 < i < n. If σ = (v0, . . . , vn) ∈ Sn, then the ith face of σ is the (n − 1)-simplex σi obtained from σ by removing the vertex vi. A simplicial complex gives rise to a topological space by means of the construc- tion known as geometric realization. In brief, one associates a point (a standard geometric 0-simplex) with each 0-simplex, a line segment (a standard geometric 1- simplex) with each 1-simplex, a filled-in triangle (a standard geometric 2-simplex) with each 2-simplex, etc., glued together along common faces. The intersection of two simplices in S, neither of which is a face of the other, is a proper subset, and hence a face, of both of them. In the geometric realization this means that the geo- metric simplices that realize the abstract simplices intersect on common faces, and hence give rise to a well-defined geometric object. A geometric n-simplex is nothing but a (n + 1)-clique, canonically realized as a geometric object. An n-simplex is In this case the said to be oriented if there is a linear ordering on its vertices. corresponding (n + 1)-clique is said to be a directed (n + 1)-clique. If S is a simplicial complex, then the union S (n) = Sn ∪ ··· ∪ S0, which is called the n-skeleton of S, is a subcomplex of S. We say that S is n-dimensional if S = S (n), and n is minimal with this property. If S is n-dimensional, and k ≤ n, then the collection Sk ∪ . . . ∪ Sn is not a subcomplex of S because it is not closed under taking subsets. However if one adds to that collection all the faces of all simplices in Sk ∪ . . . ∪ Sn, one obtains a subcomplex of S called the k-coskeleton of S, which we will denote by S(k). The computational usefulness of coskeleta will become clear when we discuss homology computation (ST1.3). Directed graphs give rise to abstract oriented simplicial complexes in a natural way. Let G = (V, E, τ ) be a directed graph. The directed flag complex associated to G is the abstract simplicial complex S = S(G), with S0 = V and whose n- simplices Sn for n ≥ 1 are (n + 1)-tuples (v0, . . . , vn), of vertices such that for each 0 ≤ i < j ≤ n, there is an edge in G from vi to vj. Notice that because of the assumptions on τ , an n-simplex in S is characterised by the (ordered) sequence (v0, . . . , vn), but not by the underlying set of vertices. For instance (v1, v2, v3) and (v2, v1, v3) are distinct 2-simplices with the same set of vertices. ST1.3. Homology, Betti numbers, and Euler characteristic. We now recall certain well known invariants of simplicial complexes arising in algebraic topol- ogy, which are preserved under a class of morphisms that is relevant in algebraic topology and that includes isomorphisms. These invariants serve to measure the 20 D(cid:32)LOTKO ET AL. “complexity” of simplicial complexes, from various topological perspectives, leading us to refer to them as metrics. Homology is an important algebraic invariant of topological spaces. In this paper we use only mod-2 simplicial homology, computationally the simplest variant of homology, which is why we choose to work with it in applications, though other types of simplicial homology may provide deeper information. We do not give a complete account of homology here, but rather an elementary description of what it is and its basic properties. Let F2 denote the field of two elements, which we denote by 0 and 1. Let S be a finite simplicial complex. Define the chain complex C∗(S, F2) to be the sequence {Cn = Cn(S, F2)}n≥0, such that Cn is the F2-vector space whose basis elements are the n-simplices σ ∈ Sn, for each n ≥ 0. In other words, the elements of Cn are formal linear combinations of n-simplices in S with coefficients in F2. For each n ≥ 0, there is a linear transformation called a differential ∂n : Cn+1 → Cn defined by ∂n(σ) = σ0 +σ1 +···+σn for every n-simplex σ, where σi is the i-th face of σ, as defined above. Having defined ∂n on the basis, one extends the definition linearly to the entire vector space Cn. The n-th Betti number βn(S) of a simplicial complex S is the F2-vector space dimension of its n-th mod 2 homology group, which is defined by Hn(S, F2) = Ker(∂n−1)/Im(∂n). ∂n as a (cid:0)Sn × Sn+1(cid:1)-matrix Dn with coefficients in F2, then one can easily Computing the Betti numbers is conceptually very easy. Let Sn denote the number of n-simplices in the simplicial complex S. If one encodes the differential compute its nullity, null(∂n), and its rank , rk(∂n), which are the F2-dimensions of the null-space and the column space of Dn, respectively. The Betti numbers of S are then a sequence of natural numbers defined by β0(S) = dimF2(C0) − rk(∂0), and βn(S) = null(∂n−1) − rk(∂n). The n-the Betti number βn counts the number of “n-dimensional holes” in the geometric realization of S. When S = S(G) is the directed flag complex of a directed graph G, both the simplices of S and these “n-dimensional holes” can be regarded as particularly important “metamotifs” [13] in the graph G. It is easy to show that the n-th Betti number of a simplicial complex S is equal to that of its (n − 1)-st coskeleton S(n−1), i.e., βn(S) = βn(S(n−1)), for all n. This observation turns out to be computationally very useful, since there is no need to store the simplices of dimension less than n − 1 that are not faces of higher dimensional simplices in order to compute βn(S). In this paper it was exactly this trick that allowed us to compute the top dimensional homology of the 42 N- complexes we worked with. Homology actually encodes far more information than what is intimated here, which can potentially be used for analyzing networks, but for the purposes of this article the description above will suffice. If S is a simplicial complex and Sn denotes the cardinality of the set of n- simplices in S, then the Euler characteristic of S is defined to be (cid:88) n≥0 χ(S) = (−1)nSn. TOPOLOGICAL ANALYSIS OF NEURAL MICROCIRCUITS 21 There is a well known, close relationship between Euler characterstic and Betti If {βn}n≥0 is the sequence of Betti numbers [5], which is expressed as follows. numbers for S, then (cid:88) n≥0 χ(S) = (−1)nβn(S). See Figure 2A for a specific example. ST1.4. Hasse Diagrams. A Hasse diagram, otherwise known as a directed acyclic graph, is a directed graph H = (V, E, τ ) with no oriented cycles. Hasse diagrams can be used to encode various combinatorial, geometric, and topologi- cal structures, such as posets and cubical complexes. Below we explain in detail how Hasse diagrams encode simplicial complexes. We include this discussion here because our computational algorithm (Algorithm 1) is based on this idea. A Hasse diagram H is said to be stratified if for each v ∈ V , every path from v to any sink has the same length. Thus in a stratified Hasse diagram the vertices are naturally partitioned into disjoint strata, where every directed path from a vertex in the k-th stratum Vk to any sink is of length k. In particular, the 0-th stratrum V0 is the set of sinks of H. Moreover, for all e ∈ E, there exists k > 0 such that τ1(e) ∈ Vk and τ2(e) ∈ Vk−1. Note that if H and H(cid:48) are isomorphic Hasse diagrams, and H is stratified, then so is H(cid:48). An orientation ς on a Hasse diagram H consists of a linear ordering <ς,v of the set Ev of edges with source v, for every vertex v of H. If H = (V, E, τ ) and H(cid:48) = (V (cid:48), E(cid:48), τ(cid:48)) are Hasse digrams equipped with orientations ς and ς(cid:48), respectively, then a morphism of oriented Hasse diagrams from (H, ς) to (H(cid:48), ς(cid:48)) is a morphism of directed graphs (α, β) : H → H(cid:48) such that for every v ∈ V , the restriction of β to a set map Ev → Eα(v) preserves the orientation, i.e, if e <ς,v e(cid:48) for some e, e(cid:48) ∈ Ev, then β(e) <ς(cid:48),α(v) β(e(cid:48)). A morphism (α, β) of oriented Hasse diagrams is an isomorphism if α and β are bijections. A stratified Hasse diagram equipped with an orientation is called admissible. Vertices in the k-th stratum of a stratified Hasse diagram H are said to be of level k. If k < n, and v, u are vertices of levels k and n respectively, then we say that v is a face of u if there is a path in H from u to v. If H is also oriented and therefore admissible, and there is a path (e1, ..., en−k) from u to v such that ei = min Eτ1(ei) for all 1 ≤ i ≤ n − k, we say that v is a front face of u. Similarly, v is a back face of u if there is a path (e1, ..., en−k) from u to v such that ei = max Eτ1(ei) for all 1 ≤ i ≤ n − k. We let Face(u) denote the set of all faces of u and Face(v)k the set of those that are of level k, while Front(u) and Back(u) denote its sets of front and back faces, respectively. See Figure S4 for an illustration of the concepts introduced above. Example 1. If G = (V, E, τ ) is a directed graph, then G can be equivalently repre- sented by an admissible Hasse diagram with level 0 vertices V , level 1 vertices E, and directed edges from each e ∈ E to its source and target. The ordering on the edges in the Hasse diagram is determined by the orientation of each edge e in G. Every simplicial complex S gives rise to an admissible Hasse diagram HS as follows. The level d vertices of HS are the d-simplices of S. There is a directed edge from each d-simplex to each of its (d − 1)-faces. The stratification on HS is thus given by dimension, and the orientation is given by the natural ordering of the faces of a simplex from front to back. See Figure S5. 22 D(cid:32)LOTKO ET AL. The Euler characteristic of a stratified Hasse diagram H = (V, E, τ ) is defined to be the integer χ(H) = (−1)kVk. (cid:88) k≥0 It is easy to see that isomorphic stratified Hasse diagrams have the same Euler characteristic. It is also straight forward to show that if H is a stratified Hasse diagram associated to a simplicial complex S, then the Euler characteristic of H coincides with that of S. ST2. Data structures and algorithms In this section we describe our basic data structures and provide a detailed overview of the algorithm that constructs the directed flag complex associated to a directed graph. We also indicate briefly how our homology computations were performed. A publicly available C++ implementation of the code will be available on http://neurotop.gforge.inria.fr/. ST2.1. Data structures. We represent an admissible Hasse diagram H cor- responding to the directed flag complex of a directed graph G = (V, E, τ ) by a reference-based data structure, using vectors to store the references to the vertices of the diagram. Each vertex v ∈ H stores the following information. (1) Ver(v): A vector of the vertices of G determining the simplex of the flag complex to which v corresponds. (2) Tar(v): A vector of references to the vertices that are targets of edges with source v. (3) Src(v): A vector of references to the vertices that are sources of edges with target v. The admissible Hasse diagram H is thus represented by an ordered set of d vectors, where d is the maximal level in H, and where the i-th vector contains the references to all level i vertices. Let Sint denote the size of integer data types, and for a given graph G = (V, E, τ ), let V and E denote the cardinalities of the corresponding sets. Each edge of the Hasse diagram is stored in two vertices of the diagram. If each reference requires Sint storage, then we require O(E · Sint) space to store all references. In addition, each vertex stores the vector of vertices in V of the simplex in the flag complex of G to which it corresponds, which requires an additional O(Sint·d) of space per vertex. The total size of a Hasse diagram is thus bounded by O((Sint · d) · V + E · Sint). In particular, the required storage space grows linearly with the number of vertices and with the number of edges. For our complexity analysis below we assume that accessing any vertex, using Tar or Src, takes O(1) time. ST2.2. Creation of the directed flag complex associated to a directed graph. We describe our algorithm that creates a directed simplicial complex given a directed graph G. The output is a Hasse diagram H, stored as the data structure described above. The identifier Ver(v) of a vertex v in H, corresponding to a simplex σ in the directed flag complex, is the vector of vertices in G that represents σ. For every level n ≥ 1 vertex v in H such that Ver(v) = [v0, . . . , vn], the algorithm additionally records a vector Uv of references to level 0 vertices u satisfying the following properties: TOPOLOGICAL ANALYSIS OF NEURAL MICROCIRCUITS 23 Algorithm 1 Directed flag complex generation. Input: A directed graph G = (V, E, τ ). Output: A Hasse diagram H representing the directed flag complex associated to G. 1: Convert G to level 0 and level 1 vertices of H (cf. Example 1). 2: for every level 1 vertex e ∈ H do 3: exist e1, e2 such that τ1(e1) = τ1(e), τ1(e2) = τ2(e) and τ2(e1) = τ2(e2) = u if then Add u to Ue; next level nodes – empty vector of references to nodes; for top–level vertex e ∈ H do for Every u ∈ Ue do 4: 5: dim = 2; 6: repeat 7: 8: 9: 10: 11: 12: 13: 14: 15: 16: 17: 18: 19: 20: 21: 22: 23: 24: until next level nodes = ∅ 25: Return H; Create a node t of a Hasse diagram; Ver(t) = [Ver(e), u]; Ut = Ue; Add e to Tar(t); Add t to Src(e); for Every bd ∈ Tar(e) do if u ∈ Ver(cbd) then Add cbd to Tar(t); Add t to Src(cbd); Ut = Ut ∩ Ucbd; Add next level nodes to H; dim = dim + 1; for Every cbd ∈ Src(bd) do Add t to next level nodes; (1) u (cid:54)= vi for all 0 ≤ i ≤ n, and (2) for every u ∈ Uv and every 0 ≤ i ≤ n, there exists an edge in G from vi to u. Finally, we assume that the graph G itself is given as an admissible Hasse dia- gram, as described in Example 1. Under these assumptions Algorithm 1 below is used to create the directed flag complex associated to G. ST2.3. Discussion of Algorithm 1. At the start of the algorithm (Line 1) only levels 0 and 1 of the Hasse diagram H, which are the same as those of the Hasse diagram representation of G itself, have been created (cf. Example 1). The for loop in the line 2 initialises the creation of the vectors Uv for level 1 vertices. For every level 1 vertex e, the vector Ue stores the references to all the 0-simplices that, together with e, will form a level 2 vertex t. The construction of level 2 vertices in H is performed during the first iteration of the repeat-until loop starting in Line 6. We analyze the generation of level 2 vertices as a generic case, since the ar- guments may clearly be generalised to higher levels. The if condition in Line 3 ensures that the vertex u will be the terminal vertex of the 2-dimensional simplex corresponding to the level 2 vertex t, created in the first iteration of the repeat-until loop (Line 6). Moreover the level 1 vertex e will correspond to a front face of the 24 D(cid:32)LOTKO ET AL. 2-simplex associated to t. Therefore, the ordering of Ver(e) can be extended to or- dering of Ver(t), as in Line 11. Thus all level 2 vertices corresponding to 2-simplices in the directed flag complex of G will be created by the algorithm. Also, since every simplex has a unique 1-dimensional front face, every 2-simplex will be created only once by this process. Notice also that the if condition in Line 3 ensures that only triangles in G consisting of three edges oriented as (v1, v2), (v2, v3), and (v1, v3) will give rise to level 2 vertices in H. It follows by induction that the analogous condition on orientations is then automatically satisfied for simplices of dimension greater than 2. To see this, fix n ≥ 2, and suppose that all simplices of dimension less than or equal to n have the desired property. Fix an n-simplex S = [v0, . . . , vn] and u ∈ US. By definition of the set US, there is an edge from vi to u for every i ∈ {0, . . . , n}. Note that u ∈ US(cid:48) for any S(cid:48) ∈ Tar(S). The previous iteration of the repeat-until loop (Line 6) created an oriented simplex from S(cid:48) together with u, of which u is the last vertex. Since the ordering of elements in S(cid:48) is a restriction of the ordering of elements in S, the ordering of a n + 1 dimensional simplex [v0, . . . , vn, u] restricted to any face yields the orientation of that face. It follows that Algorithm 1 does indeed construct a directed flag complex. We now discuss the termination of Algorithm 1. If a level n vertex v is a face of a level (n + 1) vertex w, then the last vertex u in Ver(w) is not present in Ver(v), but is listed in Uv. From Lines 12 and 21 of the algorithm it is clear that Uw ⊂ Uv and moreover that u (cid:54)∈ Uv. The cardinalities of the vectors U(−) are therefore decreasing for the newly created vertices. More precisely, for a vertex t and its faces si, there exist i such that Ut (cid:12) Usi. Level n + 1 vertices are created only if there exist a level n vertex t such that Ut (cid:54)= ∅. Since the cardinality of the U(−) decreases with each iteration of the repeat loop, the algorithm will terminate. We remark finally that the size of the directed flag complex corresponding to a given directed graph G may be exponential in the size of G. In that case, the process of creation of a complex is usually stopped at some fixed dimension n. The time complexity of Algorithm 1 is proportional to the size of the output complex H, multiplied by maximal level of a vertex in H (due to the target-source search performed in Line 15) of the algorithm. ST2.4. Homology and Betti numbers. All homology computations carried out for this paper were made with F2 coefficients, using the boundary matrix reduced by an algorithm from the PHAT [2] library. TOPOLOGICAL ANALYSIS OF NEURAL MICROCIRCUITS 25 6. Supplementary Figures Figure S1. Average number of ones in the true transmission- response matrices for different pairs of parameters (∆t1, ∆t2) in a simulation of spontaneous, in-vivo-like activity (Ca 1.2) 1251020501001251020501001.21.62.02.42.83.23.64.04.44.8Delta t2 (ms)Delta t1 (ms)log10 number of ones 26 D(cid:32)LOTKO ET AL. randomized and S2. Comparing Figure non-randomized transmission-response matrices: average number of ones in a true transmission-response (t-r) matrix divided by the average number of ones obtained when the recorded spikes were randomized before calculating the t-r matrix. Matrices were calculated from simulated spontaneous, ongoing activity with different values for ∆t1 (in different colors) and ∆t2 (along the x-axis). For each pair (∆t1, ∆t2), matrices for 20 time steps were calculated, and the mean ratio is shown. Spikes were randomized by shuffling the identities of the firing neurons, thus conserving the number of spikes in any given time step and the total number of spikes fired by each neuron. Figure S3. (A-C) Examples of directed graphs. Graphs (A) and (B) are isomorphic, where the isomorphism is given by the map sending vertex a to 1, b to 2, c to 3, and d to 4. Graphs (A) and (B) are not isomorphic to graph (C). Vertex b in graph (A) is a sink, vertex a in the same graph is a source. Graph (C) has no sources or sinks, which explains the lack of isomorphism to graphs (A) and (B). 0204060801000.80.91.0∆t1 = 1 ms∆t1 = 2 ms∆t1 = 5 ms∆t1 = 10 ms∆t1 = 20 ms∆t1 = 50 ms∆t1 = 100 ms∆t2 (ms):seno fo rebmun fo oitaRdelffuhs / lautca1.11.21.31.41.5shuffled = non-shuffledacdb12341234(A)(B)(C) TOPOLOGICAL ANALYSIS OF NEURAL MICROCIRCUITS 27 Figure S4. (A) A Hasse diagram that is not stratified, due to the edge from the vertex 1 to 5. (B) A stratified Hasse diagram, where vertices 5, 6, and 7 are the vertices of level 0, vertices 2, 3, and 4 are of level 1, and vertex 1 is of level 2. This is also an admissible Hasse diagram, where the outgoing edges are ordered from left to right. Vertex 2 is a front face of vertex 1, while vertex 3 is neither a front nor a back face of a vertex 1, and vertex 4 is back face of a vertex 1. Figure S5. Top: The geometric realization of a simplicial com- plex consisting of seven 0-simplices (labeled 1,...,7), ten 1-simplices, and four 2-simplices. The orientation on the edges is denoted by arrows, i.e., the tail of an arrow is its source vertex, while the head of an arrow is its target. Bottom: The Hasse diagram correspond- ing to the simplicial complex above. Level k vertices correspond to k-simplices of the complex and are labeled by the ordered sets of vertices that constitute the corresponding simplex. Note that, e.g., vertex 23 is a back face of a vertex 123 and a front face of a vertex 234. 123455672341(A)(B)1234512132324343545123234345135426767565767567 28 D(cid:32)LOTKO ET AL. Geometrica, Inria, Saclay, France Laboratory for Topology and Neuroscience, ´Ecole Polytechnique F´ed´erale de Lau- sanne, Lausanne, Switzerland Institute of Mathematics, University of Aberdeen, Aberdeen, UK Blue Brain Project, ´Ecole Polytechnique F´ed´erale de Lausanne, Lausanne, Switzer- land Blue Brain Project, ´Ecole Polytechnique F´ed´erale de Lausanne, Lausanne, Switzer- land Laboratory for Topology and Neuroscience, ´Ecole Polytechnique F´ed´erale de Lau- sanne, Lausanne, Switzerland Laboratory for Topology and Neuroscience, ´Ecole Polytechnique F´ed´erale de Lau- sanne, Lausanne, Switzerland Blue Brain Project, ´Ecole Polytechnique F´ed´erale de Lausanne, Lausanne, Switzer- land Blue Brain Project, ´Ecole Polytechnique F´ed´erale de Lausanne, Lausanne, Switzer- land
1605.06544
1
1605
2016-05-20T21:38:32
Inference by Reparameterization in Neural Population Codes
[ "q-bio.NC" ]
Behavioral experiments on humans and animals suggest that the brain performs probabilistic inference to interpret its environment. Here we present a new general-purpose, biologically-plausible neural implementation of approximate inference. The neural network represents uncertainty using Probabilistic Population Codes (PPCs), which are distributed neural representations that naturally encode probability distributions, and support marginalization and evidence integration in a biologically-plausible manner. By connecting multiple PPCs together as a probabilistic graphical model, we represent multivariate probability distributions. Approximate inference in graphical models can be accomplished by message-passing algorithms that disseminate local information throughout the graph. An attractive and often accurate example of such an algorithm is Loopy Belief Propagation (LBP), which uses local marginalization and evidence integration operations to perform approximate inference efficiently even for complex models. Unfortunately, a subtle feature of LBP renders it neurally implausible. However, LBP can be elegantly reformulated as a sequence of Tree-based Reparameterizations (TRP) of the graphical model. We re-express the TRP updates as a nonlinear dynamical system with both fast and slow timescales, and show that this produces a neurally plausible solution. By combining all of these ideas, we show that a network of PPCs can represent multivariate probability distributions and implement the TRP updates to perform probabilistic inference. Simulations with Gaussian graphical models demonstrate that the neural network inference quality is comparable to the direct evaluation of LBP and robust to noise, and thus provides a promising mechanism for general probabilistic inference in the population codes of the brain.
q-bio.NC
q-bio
Inference by Reparameterization in Neural Population Codes Rajkumar V. Raju Department of ECE Rice University Houston, Tx 77005 [email protected] Dept. of Neuroscience, Dept. of ECE Baylor College of Medicine, Rice University Xaq Pitkow Houston, Tx 77005 [email protected] Abstract Behavioral experiments on humans and animals suggest that the brain performs probabilistic inference to interpret its environment. Here we present a new general- purpose, biologically-plausible neural implementation of approximate inference. The neural network represents uncertainty using Probabilistic Population Codes (PPCs), which are distributed neural representations that naturally encode prob- ability distributions, and support marginalization and evidence integration in a biologically-plausible manner. By connecting multiple PPCs together as a proba- bilistic graphical model, we represent multivariate probability distributions. Ap- proximate inference in graphical models can be accomplished by message-passing algorithms that disseminate local information throughout the graph. An attractive and often accurate example of such an algorithm is Loopy Belief Propagation (LBP), which uses local marginalization and evidence integration operations to perform approximate inference efficiently even for complex models. Unfortunately, a subtle feature of LBP renders it neurally implausible. However, LBP can be elegantly reformulated as a sequence of Tree-based Reparameterizations (TRP) of the graphical model. We re-express the TRP updates as a nonlinear dynamical system with both fast and slow timescales, and show that this produces a neurally plausible solution. By combining all of these ideas, we show that a network of PPCs can represent multivariate probability distributions and implement the TRP updates to perform probabilistic inference. Simulations with Gaussian graphical models demonstrate that the neural network inference quality is comparable to the direct evaluation of LBP and robust to noise, and thus provides a promising mechanism for general probabilistic inference in the population codes of the brain. 1 Introduction In everyday life we constantly face tasks we must perform in the presence of sensory uncertainty. A natural and efficient strategy is then to use probabilistic computation. Behavioral experiments have established that humans and animals do in fact use probabilistic rules in sensory, motor and cognitive domains [1, 2, 3]. However, the implementation of such computations at the level of neural circuits is not well understood. In this work, we ask how distributed neural computations can consolidate incoming sensory in- formation and reformat it so it is accessible for many tasks. More precisely, how can the brain simultaneously infer marginal probabilities in a probabilistic model of the world? Previous efforts to model marginalization in neural networks using distributed codes invoked limiting assumptions, either treating only a small number of variables [4], allowing only binary variables [5, 6, 7], or 29th Conference on Neural Information Processing Systems (NIPS 2016), Barcelona, Spain. restricting interactions [8, 9]. Real-life tasks are more complicated and involve a large number of variables that need to be marginalized out, requiring a more general inference architecture. Here we present a distributed, nonlinear, recurrent network of neurons that performs inference about many interacting variables. There are two crucial parts to this model: the representation and the inference algorithm. We assume that brains represent probabilities over individual variables using Probabilistic Population Codes (PPCs) [10], which were derived from using Bayes' Rule on experimentally measured neural responses to sensory stimuli. Here for the first time we link multiple PPCs together to construct a large-scale graphical model. For the inference algorithm, many researchers have considered Loopy Belief Propagation (LBP) to be a simple and efficient candidate algorithm for the brain [11, 12, 13, 14, 8, 5, 7, 6]. However, we will discuss one particular feature of LBP that makes it neurally implausible. Instead, we propose that an alternative formulation of LBP known as Tree-based Reparameterization (TRP) [15], with some modifications for continuous-time operation at two timescales, is well-suited for neural implementation in population codes. We describe this network mathematically below, but the main conceptual ideas are fairly straight- forward: multiplexed patterns of activity encode statistical information about subsets of variables, and neural interactions disseminate these statistics to all other encoded variables for which they are relevant. In Section 2 we review key properties of our model of how neurons can represent probabilistic information through Probabilistic Population Codes. Section 3 reviews graphical models, Loopy Belief Propagation, and Tree-based Reparameterization. In Section 4, we merge these ingredients to model how populations of neurons can represent and perform inference on large multivariate distributions. Section 5 describes experiments to test the performance of network. We summarize and discuss our results in Section 6. 2 Probabilistic Population Codes Neural responses r vary from trial to trial, even to repeated presentations of the same stimulus x. This variability can be expressed as the likelihood function p(rx). Experimental data from several brain areas responding to simple stimuli suggests that this variability often belongs to the exponential family of distributions with linear sufficient statistics [10, 16, 17, 4, 18]: p(rx) = φ(r) exp(h(x) · r), example, if the posterior distribution is Gaussian (Figure 1), then p(xr) ∝ exp(cid:0)− 1 (1) where h(x) depends on the stimulus-dependent mean and fluctuations of the neuronal response and φ(r) is independent of the stimulus. For a conjugate prior p(x), the posterior distribution will also have this general form, p(xr) ∝ exp(h(x) · r). This neural code is known as a linear PPC: it is a Probabilistic Population Code because the population activity collectively encodes the stimulus probability, and it is linear because the log-likelihood is linear in r. In this paper, we assume responses are drawn from this family, although incorporation of more general PPCs with nonlinear sufficient statistics T(r) is possible: p(rx) ∝ exp(h(x) · T(r)). An important property of linear PPCs, central to this work, is that different projections of the population activity encode the natural parameters of the underlying posterior distribution. For with a · r and b · r encoding the linear and quadratic natural parameters of the posterior. These projections are related to the expectation parameters, the mean and variance, by µ = b·r a·r and σ2 = 1 a·r. A second important property of linear PPCs is that the variance of the encoded distribution is inversely proportional to the overall amplitude of the neural activity. Intuitively, this means that more spikes means more certainty (Figure 1). The most fundamental probabilistic operations are the product rule and the sum rule. Linear PPCs can perform both of these operations while maintaining a consistent representation [4], a useful feature for constructing a model of canonical computation. For a log-linear probability code like linear PPCs, the product rule corresponds to weighted summation of neural activities: p(xr1, r2) ∝ p(xr1)p(xr2) ⇐⇒ r3 = A1r1 + A2r2. In contrast, to use the sum rule to marginalize out variables, linear PPCs require nonlinear transformations of population activity. Specifically, a quadratic nonlinearity with divisive normalization performs near-optimal marginalization in linear PPCs [4]. Quadratic interactions arise naturally through coincidence detection, and divisive normalization is a nonlinear inhibitory effect widely observed in neural circuits [19, 20, 21]. Alternatively, near-optimal 2 x2a · r + xb · r(cid:1), 2 marginalizations on PPCs can also be performed by more general nonlinear transformations [22]. In sum, PPCs provide a biologically compatible representation of probabilistic information. Figure 1: Key properties of linear PPCs. (A) Two single trial population responses for a particular stimulus, with low and high amplitudes (blue and red). The two projections a · r and b · r encode the natural parameters of the posterior. (B) Corresponding posteriors over stimulus variables determined by the responses in panel A. The gain or overall amplitude of the population code is inversely proportional to the variance of the posterior distribution. 3 Inference by Tree-based Reparameterization 3.1 Graphical Models (cid:81) To generalize PPCs, we need to represent the joint probability distribution of many variables. A natural way to represent multivariate distributions is with probabilistic graphical models. In this work, we use the formalism of factor graphs, a type of bipartite graph in which nodes representing variables are connected to other nodes called factors representing interactions between 'cliques' or sets of variables (Figure 2A). The joint probability over all variables can then be represented as a product over cliques, p(x) = 1 c∈C ψc(xc), where ψc(xc) are nonnegative compatibility functions on the set of variables xc = {xcc ∈ C} in the clique, and Z is a normalization constant. The distribution of Z interest will be a posterior distribution p(xr) that depends on neural responses r. Since the inference algorithm we present is unchanged with this conditioning, for notational convenience we suppress this dependence on r. In this paper, we focus on pairwise interactions, although our main framework generalizes naturally to richer, higher-order interactions. In a pairwise model, we allow singleton factors ψs for variable nodes s in a set of vertices V , and pairwise interaction factors ψst for pairs (s, t) in the set of edges E that connect those vertices. The joint distribution is then (cid:89) s∈V p(x) = 1 Z (cid:89) (s,t)∈E ψs(xs) ψst(xs, xt) (2) 3.2 Belief Propagation and its neural plausibility variable, ps(xs) =(cid:82) x\xs The inference problem of interest in this work is to compute the marginal distribution for each p(x) d(x\xs). This task is generally intractable. However, the factorization structure of the distribution can be used to perform inference efficiently, either exactly in the case of tree graphs, or approximately for graphs with cycles. One such algorithm is called Belief Propagation (BP) [11]. BP iteratively passes information along the graph in the form of messages mst(xt) from node s to t, using only local computations that summarize the relevant aspects of other messages upstream in the graph: mn+1 st (xt) = dxs ψs(xs)ψst(xs, xt) mn us(xs) bs(xs) ∝ ψs mus(xs) (3) (cid:89) u∈N (s) (cid:90) xs (cid:89) u∈N (s)\t where n is the time or iteration number, and N (s) is the set of neighbors of node s on the graph. The estimated marginal, called the 'belief' bs(xs) at a node s, is proportional to the local evidence at that node ψs(xs) and all the messages coming into node s. Similarly, the messages themselves are 3 ria.rb.rb.r=a.rBANeuronindex ip(xr)xNeuralresponsePosterior1σ=a.r determined self-consistently by combining incoming messages -- except for the previous message from the target node t. This message exclusion is critical because it prevents evidence previously passed by the target node from being counted as if it were new evidence. This exclusion only prevents overcounting on a tree graph, and is unable to prevent overcounting of evidence passed around loops. For this reason, BP is exact for trees, but only approximate for general, loopy graphs. If we use this algorithm anyway, it is called 'Loopy' Belief Propagation (LBP), and it often has quite good performance [12]. Multiple researchers have been intrigued by the possibility that the brain may perform LBP [13, 14, 5, 8, 7, 6], since it gives "a principled framework for propagating, in parallel, information and uncertainty between nodes in a network" [12]. Despite the conceptual appeal of LBP, it is important to get certain details correct: in an inference algorithm described by nonlinear dynamics, deviations from ideal behavior could in principle lead to very different outcomes. One critically important detail is that each node must send different messages to different targets to prevent overcounting. This exclusion can render LBP neurally implausible, because neurons cannot readily send different output signals to many different target neurons. Some past work simply ignores the problem [5, 7]; the resultant overcounting destroys much of the inferential power of LBP, often performing worse than more naïve algorithms like mean-field inference. One better option is to use different readouts of population activity for different targets [6], but this approach is inefficient because it requires many readout populations for messages that differ only slightly, and requires separate optimization for each possible target. Other efforts have avoided the problem entirely by performing only unidirectional inference of low-dimensional variables that evolve over time [14]. Appealingly, one can circumvent all of these difficulties by using an alternative formulation of LBP known as Tree-based Reparameterization (TRP). 3.3 Tree-based Reparameterization Insightful work by Wainwright, Jakkola, and Willsky [15] revealed that belief propagation can be understood as a convenient way of refactorizing a joint probability distribution, according to approximations of local marginal probabilities. For pairwise interactions, this can be written as (cid:89) s∈V (cid:89) (s,t)∈E p(x) = 1 Z ψs(xs) ψst(xs, xt) = (cid:89) (cid:89) s∈V Ts(xs) (s,t)∈E Tst(xs, xt) Ts(xs)Tt(xt) (4) where Ts(xs) is a so-called 'pseudomarginal' distribution of xs and Tst(xs, xt) is a joint pseu- domarginal over xs and xt (Figure 2A -- B), where Ts and Tst are the outcome of Loopy Belief Propagation. The name pseudomarginal comes from the fact that these quantities are always locally consistent with being marginal distributions, but they are only globally consistent with the true marginals when the graphical model is tree-structured. These pseudomarginals can be constructed iteratively as the true marginals of a different joint distribution pτ (x) on an isolated tree-structured subgraph τ. Compatibility functions ψ from factors remaining outside of the subgraph are collected in a residual term rτ (x). This regrouping leaves the joint distribution unchanged: (5) The factors of pτ are then rearranged by computing the true marginals on its subgraph τ, again preserving the joint distribution. In subsequent updates, we iteratively refactorize using the marginals of pτ along different tree subgraphs τ (Figure 2C). p(x) = pτ (x)rτ (x) Figure 2: Visualization of tree reparameterization. (A) A probability distribution is specified by factors {ψs, ψst} on a tree graph. (B) An alternative parameterization of the same distribution in terms of the marginals {Ts, Tst}. (C) Two TRP updates for a 3× 3 nearest-neighbor grid of variables. 4 p(x)=pi(x)ri(x)p(x)=pj(x)rj(x)x1x2x3x1x2x3AOriginalIteration iIteration jTree reparameterizedBC Typical LBP can be interpreted as a sequence of local reparameterizations over just two neighboring nodes and their corresponding edge [15]. Pseudomarginals are initialized at time n = 0 using the s (xs) ∝ ψs(xs) and T 0 st(xs, xt) ∝ ψs(xs)ψt(xt)ψst(xs, xt). At iteration n + 1, original factors: T 0 the node and edge pseudomarginals are computed by exactly marginalizing the distribution built from (cid:89) previous pseudomarginals at iteration n: (cid:1) T n+1 Notice that, unlike the original form of LBP, operations on graph neighborhoods(cid:81) (cid:1)(cid:0)(cid:82) T n (cid:0)(cid:82) T n st ∝ T n+1 s u∈N (s) T n su dxu ∝ T n 1 T n s T n+1 t T n+1 s st dxs st dxt (cid:90) T n st (6) s u∈N (s) do not differentiate between targets. 4 Neural implementation of TRP updates 4.1 Updating natural parameters TRP's operation only requires updating pseudomarginals, in place, using local information. These are appealing properties for a candidate brain algorithm. This representation is also nicely compatible with the structure of PPCs: different projections of the neural activity encode the natural parameters of an exponential family distribution. It is thus useful to express the pseudomarginals and the TRP inference algorithm using vectors of sufficient statistics φc(xc) and natural parameters θn c for each c · φc(xc)). For a model with at most pairwise interactions, the TRP (cid:88) clique: T n updates (6) can be expressed in terms of these natural parameters as c (xc) = exp (θn st = θn θn+1 st + Qsθn+1 s + Qtθn+1 t + gE(θn st) (7) s = (1 − ds)θn θn+1 s + gV (θn su) u∈N (s) where ds is the number of neighbors of s, and Qs, gV and gE are matrices and nonlinear functions (for vertices V and edges E) that are determined by the particular graphical model (see below). Since the natural parameters reflect log-probabilities, the product rule for probabilities becomes a linear sum in θ, while the sum rule for probabilities must be implemented by nonlinear operations g on θ. In the concrete case of a Gaussian graphical model, the joint distribution is given by p(x) ∝ exp (− 1 2 x(cid:62)Ax + b(cid:62)x), where A and b are the natural parameters, and the linear and quadratic functions x and xx(cid:62) are the sufficient statistics. When we reparameterize this distribution by pseudomarginals, we again have linear and quadratic sufficient statistics: two for each node, φs = (− 1 t , xs, xt)(cid:62). Each of these vectors of sufficient statistics has its own vector of natural parameters, θs and θst. To approximate the marginal probabilities, the TRP algorithm initializes the pseudomarginals to (cid:62). To update θ, we must extract the matrices Q θ0 s = (Ass, bs) and nonlinear functions g that recover the univariate marginal distribution of a bivariate gaussian Tst. s, xs)(cid:62), and five for each edge, φst = (− 1 st = (Ass, Ast, Att, bs, bt) s, xsxt, − 1 (cid:62) and θ0 2 x2 2 x2 2 x2 For Tst(xs, xt) ∝ exp(cid:0)− 1 (cid:90) Ts(xs) = s − θ2xsxt − 1 2 θ1x2 dxt Tst(xs, xt) ∝ exp 2 θ3x2 (cid:1), this marginal is (cid:19) θ4θ3 − θ2θ5 xs + Using this, we can determine the form of the weight matrices and the nonlinear functions in the TRP updates (7). (8) (9) (10) 2 t + θ4xs + θ5xt x2 s 2 (cid:18) − θ1θ3 − θ2 (cid:18)0 θ3 Qt = θ3 0 0 0 1 (cid:19)(cid:62) (cid:33)(cid:62) (cid:33)(cid:62) 0 0 1 0 0 3;su − θn θn 4;suθn 2;suθn 5;su , , θn 3;su θ2;stθ5;st θ3;st , θ2;stθ4;st θ1;st 0 0 0 0 θn 1;suθn 0 (cid:18)1 (cid:32) (cid:32)(cid:0)θn 0 0 0 1 (cid:19)(cid:62) 3;su −(cid:0)θn (cid:0)θn (cid:1)2 θn 3;su 2;su (cid:1)2 (cid:1)2 2;st θ3;st , 0, 2;st θ1;st Qs = gV (θn su) = gE(θn st) = where θi;st is the ith element of θst. Notice that these nonlinear functions are all quadratic functions with a linear divisive normalization. 5 4.2 Separation of Time Scales for TRP Updates An important feature of the TRP updates is that they circumvent the 'message exclusion' problem of LBP. The TRP update for the singleton terms, (6) and (7), includes contributions from all the neighbors of a given node. There is no free lunch, however, and the price is that the updates at time n + 1 depend on previous pseudomarginals at two different times, n and n + 1. The latter update is therefore instantaneous information transmission, which is not biologically feasible. To overcome this limitation, we observe that the brain can use fast and slow timescales τfast (cid:28) τslow instead of instant and delayed signals. We convert the update equations to continuous time, and θ = introduce auxiliary variables θ which are lowpass-filtered versions of θ on a slow timescale: τslow −θ + θ. The nonlinear dynamics of (7) are then updated on a faster timescale τfast according to (cid:88) u∈N (s) θs = −ds τfast θs + gV (θsu) τfast θst = Qsθs + Qtθt + gE(θst) (11) where the nonlinear terms g depend only on the slower, delayed activity θ. By concatenating these two sets of parameters, Θ = (θ, θ), we obtain a coupled multidimensional dynamical system which represents the approximation to the TRP iterations: (12) Here the weight matrix W and the nonlinear function G inherit their structure from the discrete-time updates and the lowpass filtering at the fast and slow timescales. Θ = W Θ + G(Θ) 4.3 Network Architecture To complete our neural inference network, we now embed the nonlinear dynamics (12) into the population activity r. Since different projections of the neural activity in a linear PPC encode natural parameters of the underlying distribution, we map neural activity onto Θ by r = U Θ (13) where U is a rectangular Nr × NΘ embedding matrix that projects the natural parameters and their low-pass versions into the neural response space. These parameters can be decoded from the neural activity as Θ = U +r, where U + is the pseudoinverse of U. Applying this basis transformation to (12), we have r = U Θ = U (W Θ + G(Θ)) = U W U +r + UG(U+r). We then obtain the general form of the updates for the neural activity r = WLr + GN L(r) (14) where WLr = U W U +r and GN L(r) = UG(U +r) correspond to the linear and nonlinear computa- tional components that integrate and marginalize evidence, respectively. The nonlinear function on r inherits the structure needed for the natural parameters, such as a quadratic polynomial with a divisive normalization used in low-dimensional Gaussian marginalization problems [4], but now expanded to high-dimensional graphical models. Figure 3 depicts the network architecture for the simple graphical model from Figure 2A, both when there are distinct neural subpopulations for each factor (Figure 3A), and when the variables are fully multiplexed across the entire neural population (Figure 3B). These simple, biologically-plausible neural dynamics (14) represent a powerful, nonlinear, fully-recurrent network of PPCs which implements the TRP update equations on an underlying graphical model. 5 Experiments We evaluate the performance of our neural network on a set of small Gaussian graphical models with up to 400 interacting variables. The networks time constants were set to have a ratio of τslow/τfast = 20. Figure 4 shows the neural population dynamics as the network performs inference, along with the temporal evolution of the corresponding node and pairwise means and covariances. The neural activity exhibits a complicated timecourse, and reflects a combination of many natural parameters changing simultaneously during inference. This type of behavior is seen in neural activity recorded from behaving animals [23, 24, 25]. 6 Figure 3: Distributed, nonlinear, recurrent network of neurons that performs probabilistic inference on a graphical model. (A) This simple case uses distinct subpopulations of neurons to represent different factors in the example model in Figure 2A. (B) A cartoon shows how the same distribution can be represented as distinct projections of the distributed neural activity, instead of as distinct populations. In both cases, since the neural activities encode log-probabilities, linear connections are responsible for integrating evidence while nonlinear connections perform marginalization. Figure 4: Dynamics of neural population activity (left) and the expectation parameters of the posterior distribution that the population encodes (right) for one trial of the tree model in Figure 2A. Figure 5 shows that our recurrent neural network accurately infers the marginal probabilities, and reaches almost the same conclusions as loopy belief propagation. The data points are obtained from multiple simulations with different graph topologies, including graphs with many loops. Figure 6 verifies that the network is robust to noise even when there are few neurons per inferred parameter; adding more neurons improves performance since the noise can be averaged away. Figure 5: Inference performance of our neural network (blue) and standard loopy belief propa- gation (red) for a variety of graph topologies, including square grids up to 20 × 20 and densely connected graphs with up to 25 variables. The expectation parameters (means, covariances) of the pseudomarginals closely match the corresponding parameters for the true marginals. 6 Conclusion We have shown how a biologically-plausible nonlinear recurrent network of neurons can repre- sent a multivariate probability distribution using population codes, and can perform inference by reparameterizing the joint distribution to obtain approximate marginal probabilities. 7 linearconnectionssingletonprojectionspairwiseprojectionsnonlinearconnectionslinearconnectionssingletonpopulationspairwisepopulationsnonlinearconnectionsr12r23r1r2r3ABminmaxTimeNeural activity rTimeTimeInferred expectationparametersMeansCovariancesTrue meanInferred meanNeural networkLBPTrue varianceInferred varianceTrue covarianceInferred covariance Figure 6: Network performance is robust to noise, and improves with more neurons. (A) Neural activity performing inference on a 5 × 5 square grid, in the presence of independent spatiotemporal Gaussian noise of standard deviation 0.1 times the standard deviation of each signal. (B) Expectation parameters (means, variances) of the node pseudomarginals closely match the corresponding parame- ters for the true marginals, despite the noise. Results are shown for one or five neurons per parameter in the graphical model, and for no noise (i.e. infinitely many neurons). Our network model has desirable properties beyond those lauded features of belief propagation. First, it allows for a thoroughly distributed population code, with many neurons encoding each variable and many variables encoded by each neuron. This is consistent with neural recordings in which many task-relevant features are multiplexed across a neural population [23, 24, 25]. Second, the network performs inference in place, without using a distinct neural representation for messages, and avoids the biological implausibility associated with sending different messages about every variable to different targets. This virtue comes from exchanging multiple messages for multiple timescales. It is noteworthy that allowing two timescales prevents overcounting of evidence on loops of length two (target to source to target). This suggests a novel role of memory in static inference problems: a longer memory could be used to discount past information sent at more distant times, thus avoiding the overcounting of evidence that arises from loops of length three and greater. It may therefore be possible to develop reparameterization algorithms with all the convenient properties of LBP but with improved performance on loopy graphs. Previous results show that the quadratic nonlinearity with divisive normalization is convenient and biologically plausible interpretable, but this precise form is not necessary: other pointwise neuronal nonlinearities are capable of producing the same high-quality marginalizations in PPCs [22]. In a distributed code, the precise nonlinear form at the neuronal scale is not important as long as the effect on the parameters is the same. More generally, however, different nonlinear computations on the parameters implement different approximate inference algorithms. The distinct behaviors of such algorithms as mean-field inference, generalized belief propagation, and others arise from differences in their nonlinear transformations. Even Gibbs sampling can be described as a noisy nonlinear message-passing algorithm. Although LBP and its generalizations have strong appeal, we doubt the brain will use this algorithm exactly. The real nonlinear functions in the brain may implement even smarter algorithms. To identify the brain's algorithm, it may be more revealing to measure how information is represented and transformed in a low-dimensional latent space embedded in the high-dimensional neural responses than to examine each neuronal nonlinearity in isolation. The present work is directed toward this challenge of understanding computation in this latent space. It provides a concrete example showing how distributed nonlinear computation can be distinct from localized neural computations. Learning this computation from data will be a key challenge for neuroscience. In future work we aim to recover the latent computations of our network from artificial neural recordings generated by the model. Successful model recovery would encourage us to apply these methods to large-scale neural recordings to uncover key properties of the brain's distributed nonlinear computations. Author contributions XP conceived the study. RR and XP derived the equations. RR implemented the computer simulations. RR and XP analyzed the results and wrote the paper. 8 minmaxTrue parametersNneuronsBAno noiseNeural activity rInferred parametersMeanVarianceTimeNparams= 1NneuronsNparams= 5 Acknowledgments: XP and RR were supported by a grant from the McNair Foundation and by the Intelligence Advanced Research Projects Activity (IARPA) via Department of Interior/Interior Business Center (DoI/IBC) contract number D16PC00003.1 References [1] Knill DC, Richards W (1996) Perception as Bayesian inference. Cambridge University Press. [2] Doya K (2007) Bayesian brain: Probabilistic approaches to neural coding. MIT press. [3] Pouget A, Beck JM, Ma WJ, Latham PE (2013) Probabilistic brains: knowns and unknowns. Nature neuroscience 16: 1170 -- 1178. [4] Beck JM, Latham PE, Pouget A (2011) Marginalization in neural circuits with divisive normalization. The Journal of neuroscience 31: 15310 -- 15319. [5] Ott T, Stoop R (2006) The neurodynamics of belief propagation on binary markov random fields. In: Advances in neural information processing systems. pp. 1057 -- 1064. [6] Steimer A, Maass W, Douglas R (2009) Belief propagation in networks of spiking neurons. Neural Computation 21: 2502 -- 2523. [7] Litvak S, Ullman S (2009) Cortical circuitry implementing graphical models. Neural computation 21: 3010 -- 3056. [8] George D, Hawkins J (2009) Towards a mathematical theory of cortical micro-circuits. PLoS Comput Biol 5: e1000532. [9] Grabska-Barwinska A, Beck J, Pouget A, Latham P (2013) Demixing odors-fast inference in olfaction. In: Advances in Neural Information Processing Systems. pp. 1968 -- 1976. [10] Ma WJ, Beck JM, Latham PE, Pouget A (2006) Bayesian inference with probabilistic population codes. Nature neuroscience 9: 1432 -- 1438. [11] Pearl J (1988) Probabilistic reasoning in intelligent systems: networks of plausible inference. Morgan Kaufmann. [12] Yedidia JS, Freeman WT, Weiss Y (2003) Understanding belief propagation and its generalizations. Exploring artificial intelligence in the new millennium 8: 236 -- 239. [13] Lee TS, Mumford D (2003) Hierarchical bayesian inference in the visual cortex. JOSA A 20: 1434 -- 1448. [14] Rao RP (2004) Hierarchical bayesian inference in networks of spiking neurons. In: Advances in neural information processing systems. pp. 1113 -- 1120. [15] Wainwright MJ, Jaakkola TS, Willsky AS (2003) Tree-based reparameterization framework for analysis of sum-product and related algorithms. Information Theory, IEEE Transactions on 49: 1120 -- 1146. [16] Jazayeri M, Movshon JA (2006) Optimal representation of sensory information by neural populations. Nature neuroscience 9: 690 -- 696. [17] Beck JM, Ma WJ, Kiani R, Hanks T, Churchland AK, Roitman J, Shadlen MN, et al. (2008) Probabilistic population codes for bayesian decision making. Neuron 60: 1142 -- 1152. [18] Graf AB, Kohn A, Jazayeri M, Movshon JA (2011) Decoding the activity of neuronal populations in macaque primary visual cortex. Nature neuroscience 14: 239 -- 245. [19] Heeger DJ (1992) Normalization of cell responses in cat striate cortex. Visual neuroscience 9: 181 -- 197. [20] Carandini M, Heeger DJ (2012) Normalization as a canonical neural computation. Nature Reviews Neuroscience 13: 51 -- 62. [21] Rubin DB, Van Hooser SD, Miller KD (2015) The stabilized supralinear network: A unifying circuit motif underlying multi-input integration in sensory cortex. Neuron 85: 402 -- 417. [22] Vasudeva Raju R, Pitkow X (2015) Marginalization in random nonlinear neural networks. In: COSYNE Abstract. [23] Hayden BY, Platt ML (2010) Neurons in anterior cingulate cortex multiplex information about reward and action. The Journal of Neuroscience 30: 3339 -- 3346. [24] Rigotti M, Barak O, Warden MR, Wang XJ, Daw ND, Miller EK, Fusi S (2013) The importance of mixed selectivity in complex cognitive tasks. Nature 497: 585 -- 590. [25] Raposo D, Kaufman MT, Churchland AK (2014) A category-free neural population supports evolving demands during decision-making. Nature neuroscience 17: 1784 -- 1792. 1The U.S. Government is authorized to reproduce and distribute reprints for Governmental purposes notwith- standing any copyright annotation thereon. Disclaimer: The views and conclusions contained herein are those of the authors and should not be interpreted as necessarily representing the official policies or endorsements, either expressed or implied, of IARPA, DoI/IBC, or the U.S. Government. 9
1506.06717
2
1506
2015-11-22T11:50:07
How we move is universal: scaling in the average shape of human activity
[ "q-bio.NC" ]
Human motor activity is constrained by the rhythmicity of the 24 hours circadian cycle, including the usual 12-15 hours sleep-wake cycle. However, activity fluctuations also appear over a wide range of temporal scales, from days to a few seconds, resulting from the concatenation of a myriad of individual smaller motor events. Furthermore, individuals present different propensity to wakefulness and thus to motor activity throughout the circadian cycle. Are activity fluctuations across temporal scales intrinsically different, or is there a universal description encompassing them? Is this description also universal across individuals, considering the aforementioned variability? Here we establish the presence of universality in motor activity fluctuations based on the empirical study of a month of continuous wristwatch accelerometer recordings. We study the scaling of average fluctuations across temporal scales and determine a universal law characterized by critical exponents $\alpha$, $\tau$ and $1/{\mu}$. Results are highly reminiscent of the universality described for the average shape of avalanches in systems exhibiting crackling noise. Beyond its theoretical relevance, the present results can be important for developing objective markers of healthy as well as pathological human motor behavior.
q-bio.NC
q-bio
Papers in Physics, vol. 7, art. 070017 (2015) Received: 18 October 2015, Accepted: 10 November 2015 Edited by: E. Mizraji Reviewed by: J. Lin, Department of Physics, Washington College, Maryland, USA. Licence: Creative Commons Attribution 3.0 DOI: http://dx.doi.org/10.4279/PIP.070017 www.papersinphysics.org ISSN 1852-4249 How we move is universal: Scaling in the average shape of human activity Dante R. Chialvo,1 Ana Mar´ıa Gonzalez Torrado,2 Ewa Gudowska-Nowak,3 Jeremi K. Ochab,4 Pedro Montoya,2 Maciej A. Nowak,3, 4 Enzo Tagliazucchi5 Human motor activity is constrained by the rhythmicity of the 24 hours circadian cycle, including the usual 12-15 hours sleep-wake cycle. However, activity fluctuations also ap- pear over a wide range of temporal scales, from days to a few seconds, resulting from the concatenation of a myriad of individual smaller motor events. Furthermore, individuals present different propensity to wakefulness and thus to motor activity throughout the cir- cadian cycle. Are activity fluctuations across temporal scales intrinsically different, or is there a universal description encompassing them? Is this description also universal across individuals, considering the aforementioned variability? Here we establish the presence of universality in motor activity fluctuations based on the empirical study of a month of con- tinuous wristwatch accelerometer recordings. We study the scaling of average fluctuations across temporal scales and determine a universal law characterized by critical exponents α, τ and 1/µ. Results are highly reminiscent of the universality described for the average shape of avalanches in systems exhibiting crackling noise. Beyond its theoretical relevance, the present results can be important for developing objective markers of healthy as well as pathological human motor behavior. 1 Consejo Nacional de Investigaciones Cient´ıficas y Tec- nol´ogicas (CONICET), Rivadavia 1917, Buenos Aires, Argentina. 2 Institut Universitari d'Investigacions en Ci`encies de la Salut (IUNICS) & Universitat de les Illes Balears (UIB), Palma de Mallorca, Spain. 3 M. Kac Complex Systems Research Center and M. Smoluchowski Institute of Physics, Jagiellonian Univer- sity, Krak´ow, Poland. 4 Biocomplexity Department, Ma(cid:32)lopolska Center of Biotechnology, Jagiellonian University, Krak´ow, Poland. 5 Institute for Medical Psychology, Christian Albrechts University, Kiel, Germany. I. Introduction The most obvious periodicity of human (as well as animal) motor activity is the circadian twenty four hours modulation. However, smaller fluctuations are evident on a wide range of temporal scales, from days to a few seconds. Data shows that the activity evolves in bursts of all sizes and durations which are known to be scale-invariant [1–8] regardless of the origins and intended consequences of such activity. Despite the variety of results, the mechanisms un- derlying the scale-invariant behavior of motor ac- tivity remain to be elucidated. Considering the in- termittent nature of human motor activity - com- prising brief activity excursions separated by peri- ods of quiescence - a natural approach would be to study the average shape of the events, following re- cent results [9–12] which show that for a large class of processes, the average shape is a scaling function determined mostly by the temporal correlations of 070017-1 5 1 0 2 v o N 2 2 ] . C N o i b - q [ 2 v 7 1 7 6 0 . 6 0 5 1 : v i X r a Papers in Physics, vol. 7, art. 070017 (2015) / D. Chialvo et al. the process and its nonlinearities [13]. In the present work, long time series of human motor activity are analyzed, recorded via wrist- watch accelerometer, lasting approximately one month. We establish first the presence of truncated scale-invariance in the distribution of the durations of the events as well as in its power spectral den- sity, as described previously in similar type of data. Afterwards, we uncover the average shape of the bursts of activity and derive the scaling function and its associated exponents. Finally, we discuss the origins of such scaling and some possible appli- cations. II. Materials and methods The recordings analyzed were part of a larger study and included six healthy, non-smokers, drug-free volunteers (mean age 50.1 years, S.D. = 6.8). The study was approved by the Bioethics Commission of the University of Isles Baleares (Spain). Par- ticipants were informed about the procedures and goals of the study, and provided their written con- sent. After determining their handedness, each subject was provided with a wristwatch-sized activ- ity recorder (Actiwatch from Mini- Mitter Co., OR, USA) measuring acceleration changes in the fore- arm in any plane. Each data point of activity corre- sponded to the number of zero crossings in acceler- ation larger than 0.01 G (sampled at 32 Hz and in- tegrated over a 30-second window length). Records of several thousands of data points were kept in the device's internal memory until being downloaded to a personal computer every week. Subjects wore the device in their non-dominant arm continuously for up to several weeks (mean 28.1 days, S.D.= 4.). Af- ter careful visual inspection of the data to exclude sets with gaps (due to subject non-compliance), a combined total of 280 days of data was available for further analysis. III. Results Figure 1: Example data set, distribution of suc- cessive increments and their spectral power. Panel A: Time series of activity x(n) recorded continu- ously from a subject during a month. Individual traces correspond to consecutive days. The top subpanel depicts daily activity averaged over the entire month. Panel B: Time series of successive increments I(n) = x(n + 1) − x(n) (normalized by its SD) for the same data. Panel C: Probability density distribution of the time series of succes- sive increments I(n) (continuous line), exhibiting exponential tails (compare with the dotted line, a Gaussian of the same variance). Panel D: Power spectral density (black line) of the time series of successive increments I(n) of panel B. This is scale invariant S(f ) ∼ f γ with γ = 0.9 (dashed line). In contrast, for the randomly shuffled increments, the serial correlations vanish and a flat spectral density is obtained (red). For ease of presentation, we will use recordings from a single subject to describe the main results. Nev- ertheless, results are robust as well as similar for the entire group of subjects in the study. A typical recording is presented in Fig. 1. Panel A shows a full month of continuously recorded activity from this subject, who is particularly regular in her daily routines. The subject wakes up with the alarm clock at 6:45 a.m. on week days and has lunch fol- lowed by a short nap each day (between 2:00 p.m. and 4:00 p.m. Panel B displays the time series of the successive increments of the signal x(n), defined 070017-2 68101214161820220246Mon12/14Tue12/15Wed12/16Thu12/17Fri12/18Sat12/19Sun12/20Mon12/21Tue12/22Wed12/23Thu12/24Fri12/25Sat12/26Sun12/27Mon12/28Tue12/29Wed12/30Thu12/31Fri01/01Sat01/02Sun01/03Mon01/04Tue01/05Wed01/06Thu01/07Fri01/08Sat01/09Sun01/10Mon01/11Tue01/12HourDate6810121416182022024600.51Activity (a.u)07142128−505Time(days)I(n)/S.D.−505−4−3−2−10I/S.D.log10P(I)/P(0)−6−5−4−3−2−6−4−20log10f(sec−1)log10S(f)BCDA1 day1 hour1 min. Papers in Physics, vol. 7, art. 070017 (2015) / D. Chialvo et al. S(f ) ∼ f β [1, 2]. Because this type of processes are likely to be non-stationary, it is best to esti- mate the exponents of the spectral density by doing the calculations over the time series of successive increments, whose density distribution is station- ary. For instance, for Brownian motion (which is summed white noise), the power spectrum decays S(f ) ∼ f β with β = −2 and for white noise β = 0; the summed time series has an exponent +2 larger than the non-summed time series. As discussed in [14], this can be generalized for all self-affine pro- cesses: summing a self-affine time series shifts the theoretical power-spectral density exponent by +2, and the reverse process is also true: the differences in consecutive values (the "first differences") of a Brownian motion result in white noise, thus tak- ing the first differences shifts the theoretical power- spectral density exponent β by −2. In our case, the exponent obtained for the time series of successive increments I(n) was γ = 0.9. Thus, the exponent of the raw data is β = γ − 2 = −1.1 [14]. For com- parison, the spectral densities of the actual signal and of a surrogate obtained after randomly shuf- fling the increments are jointly displayed in Panel D of Fig. 1. To further study the time series from the perspec- tive of individual bursts of activity, we introduce the definition of an event. We consider the time series of activity x(n) and select a threshold value U to be vanishingly small. An event is defined by the consecutive points starting when x(n) > U and ending when x(n) < U . This is equivalent to the definition of avalanches in other contexts [9,15]. In the following part, we will be concerned primarily with the statistics of event lifetimes T , as well as of their average size S and shape. In all subjects, we found that the distributions of event durations and sizes (defined by the area, i.e., the integral of the signal corresponding to the individual events) can be well described, for relatively small values, by a power-law (Fig. 2, Panel A). In contrast, the dis- tribution of waiting times between events demon- strated an exponential decay. In addition to the scale invariance, we found that the longer an event lasted, the stronger the motor activity executed by the subject. The plot of average event size (cid:104)S(cid:105) as a function of duration T follows a power-law (for small values of T ) described by (cid:104)S(cid:105)(T ) = T µ+1 with µ + 1 = 1.59. The exponents in this power-law are robust across subjects and to changes of threshold Figure 2: Scaling of activity events in a single subject (same dataset as in Fig. 1). Panel A: The complementary cumulative distribution func- tion (CCDF) for event durations (T) and sizes (S) obeys power-laws with exponents α(cid:48) = 0.70 and τ(cid:48) = 0.44, respectively (dashed lines). Note that here the densities are cumulative, thus the expo- nents of the respective PDFs are α = α(cid:48) + 1 and τ = τ(cid:48) + 1. The waiting time between events falls exponentially. Panel B: The average size of a given duration is well described (for small T) by (cid:104)S(cid:105)(T ) ∼ T µ+1 with µ + 1 = 1.59 (blue dashed line) comparable with results obtained from fitting within the scaling region (red filled symbols) giving µ + 1 = 1.61. as I(n) = x(n + 1) − x(n). The large-scale statistical features of the time se- ries presented in Fig. 1 are already well known. The density distribution of the successive incre- ments i(n) is non-Gaussian, as can be appreci- ated by a joint plot with a Gaussian distribution of the same variance (Fig. 1, Panel C). It is known that the power spectrum of the activity decays as 070017-3 102103104T102103104105<S> 100101102103104105S,Wt,T10-310-210-1100CCDFSWtTABa't'm + 1 Papers in Physics, vol. 7, art. 070017 (2015) / D. Chialvo et al. cesses in cells [17], crackling noise [11], the statis- tics of Barkhausen noise in permalloy thin films [10] and plastic deformation of metals [18]. In all these cases, the distributions obey universal functional forms: (1) (2) (3) f (S) ∼ S−τ , f (T ) ∼ T −α, (cid:104)S(cid:105)(T ) ∼ T 1/σνz, where f denotes the probability density functions of the size of the event S and its duration T , and (cid:104)S(cid:105)(T ) is the expected size for a given duration. The parameters τ , α and 1/σνz are the critical ex- ponents of the system and are expected to be inde- pendent of the details, being related to each other by the scaling relation: α − 1 τ − 1 σνz (4) 1 . = We found that the empirical exponents very closely fulfill the expression above. Using the fitting approach introduced by Clauset [19] in the scaling regions depicted in Panel A of Fig. 2, we found τ = 1.44 and α = 1.70 . Thus, from Eq. (4) a value of 1/σνz = µ + 1 = 1.59 is expected. The experimental data points are very close to this the- oretical expectation (dashed line), especially for the relatively small T values within the scaling region of Panel A (where a linear fit estimates µ + 1 = 1.61), while those for relatively larger T values (corre- sponding to the cutoff of the distributions) are a bit apart, probably due to undersampling. After repeating this analysis for all subjects in our sam- ple, the average exponents were all within 5% of the reported values. From scaling arguments, it is expected that the average shape of an event of duration T (cid:104)x(T, t)(cid:105) scales as : (cid:104)x(T, t)(cid:105) = T µfshape(t/T ). (5) Thus, the shapes of events of different durations T rescaled by µ should collapse on a single scaling function given by fshape(t/T ). Note that µ corre- sponds in this context to the wandering exponent (i.e., the mean squared displacement) of the activ- ity [13, 20]. Figure 3: Collapse of events of different duration into a single functional form. Panel A: Three ex- amples of typical events of duration T=480, 960 and 1920 sec.. Panel B: The heterogeneous events shown in Panel A can be collapsed onto the av- erage shape (dashed black line) by normalizing t to t/T and (cid:104)x(t)(cid:105) to (cid:104)x(t)(cid:105)/T µ. The inset shows the cumulative variance for a range of µ. Panel C: The average event shape, i.e., fshape(t/T ), re- covered from six data sets (thin lines). The best fit using an inverted parabola is shown as a red dashed line (µ = 0.49) as well as the one expected from the critical exponent µ = 0.59 as a dot-dashed blue line. over a reasonable range of values. This type of scaling is well known in the statisti- cal mechanics of critical phenomena [15]. Examples range from earthquakes [16] to active transport pro- Examples of this collapse are presented in Pan- els A and B of Fig. 3. Considering the number of events here averaged (in the order of N ∼ 102), the 070017-4 00.20.40.60.81 t/T051015<x> / T m0.20.40.6m0.10.150.2e00.20.40.60.81t/T00.20.40.60.81 <x> / T m<x(t/T)> / <x(t/T)>max(4(t/T)(1-t/T))0.48(4(t/T)(1-(t/T))0.59060012001800Time(sec)0150300Activity <x(t)>TABC Papers in Physics, vol. 7, art. 070017 (2015) / D. Chialvo et al. studied using this method. The best fit disagrees with the empirical functions near their peak, the latter being flatter, likely an effect related to satu- ration observed in long events. Finally, we turn to discuss simple null models. We consider two extreme cases, in both of them the raw time series are randomly shuffled to remove serial correlations. In the first case, we remove all temporal correlations by randomly reordering x(n), thus attaining a flat power spectral density. After repeating the above analysis in this surrogate data set, it becomes clear (as shown in Fig. 4) that the scale invariance is absent in all the statistics un- der study: size S, waiting time Wt and duration T of events (note that the distributions are here plotted using a logarithmic-linear scale). Results in Panel B show that µ + 1 = 1, thus µ = 0, imply- ing that there is not collapse, because with T µ = 1 in Eq. (5), the amplitude of the individual events remains invariable. To consider the second case, we need first to reorder randomly the time series of increments I(n) and then proceed to integrate the increments. Since each increment is now a random variable, the power spectral density for this surro- gate process obeys f β with β = −2 , and as shown analytically by Baldassari et al. [13], for this case µ = 1/2 and the scaling function is a semicircle. Please note that the fluctuations of human activity described here differ from a simple auto-regressive process: indeed successive increments I(n) are anti- correlated and the power spectral density corre- sponds to non-trivial power law correlations (i.e., β (cid:54)= −2). Figure 4: Scaling is absent in a null model result- ing from defining events after randomly reordering the time series x(n). Panel A: Density distribu- tions (CCDF) for event duration, size, and waiting time. All the distributions are exponential (note the logarithmic-linear scale). Panel B: The ex- pected average size for a given duration in the null model is a linear function of T (the dashed line rep- resents the fit with slope 1), therefore, µ = 0 and there is no collapse. IV. Discussion data collapse is quite satisfactory, while the value of the exponent (µ = 0.48) does not exactly match the one predicted in Eq. (4), µ = 0.59 (likely a conse- quence of insufficient sampling). To determine the generality of our results, we extended this analy- sis to six other data sets. For each data set, the value of µ was first determined. Subsequently, the x(T, t) obtained from the events were rescaled with T µ and their average computed. To account for individual differences in mean activity, shape func- tions were normalized by their mean value. The results for the six datasets are presented in Panel C of Fig. 3. They can be accurately described by an inverted parabola, as in other systems previously The present findings can be summarized by six styl- ized facts describing bursts of human activity: I) the spectral density of the time series of activity x(t) obeys a power law, with exponent β ∼ 1; II) successive increments I(n) are anti-correlated with a spectral density obeying a power law with exponent γ ∼ 1, which corresponds to a spec- tral density for the raw data f β with β ∼ −1; III) the PDF of the increments I(n) is definitely non-gaussian; IV) the PDF of duration and sizes of events obeys truncated power laws with expo- nents 1 < τ < 2 and 1 < α < 2; V) the aver- age size of the events scales with its lifetime T as 070017-5 102103T102103104<S> 020004000S,Wt,T10-410-310-210-1100CCDFSWtTAB Papers in Physics, vol. 7, art. 070017 (2015) / D. Chialvo et al. (cid:104)S(cid:105)(T ) ∼ T µ, where µ + 1 = (α − 1)/(τ − 1); VI) the time series of individual events can be appropri- ately rescaled via a transformation of its duration T and amplitude x(t) onto a unique functional shape: (cid:104)x(T, t)(cid:105) = T µfshape(t/T ). We are aware that these observations are novel only for human activity, because similar statistical regularities of avalanching activity are well known for a large variety of inanimate systems [9–12]. The rescaling of the average shape is not surprising be- cause, placed in the appropriate context, it can be traced back to Mandelbrot's study of the fractal properties of self-affine functions [21]. A curve or a time series are said to be self-affine if a transfor- mation can be found, such that rescaling their x, y coordinates by k and kµ, respectively, and the vari- ance in y is preserved (with µ = 1 corresponding to self-similarity). In that sense, the successful col- lapse of the events shape is a trivial consequence of the overall self-affinity of the x(t) time series. Thus, it is clear that the existence of the scal- ing uncovered here is not informative per se of the type of mechanism behind: scale-invariance can be constructed via different processes, ranging from critical phenomena [15] to simple stochastic auto- regressive dynamics [13, 20]. What is then the mechanism by which the above six facts are gen- erated? It seems that this question cannot be easily an- swered by the type of experiments reported here. Fluctuations of this type could have either an in- trinsic (i.e., brain-born) origin but also could be the reflection of a collective phenomena (including humans and its environment). In either case, the correlations observed seem to reject the case of in- dependent random events starting and stopping hu- man actions, because neither the distribution of the increments I(n), nor the exponents match the case of a random walk. In terms of brain-born process, it is hard to accept some of the implications of the scaling function in the activity shape. The average parabolic shape means that the very beginning of the motion activity contains information about how long the activity will last, in the same sense that the initial trajectory of a projectile predicts when and where it will land. This proposal is hardly realistic, because there is hardly a reasonable physiological argument in support of any motor planning for the length of time we are observing (∼ 103 secs). In terms of collective processes, the results here sug- gest that the interaction with other humans could determine when and where, on the average, we start and stop moving. Despite our current relative ignorance, a possi- bility that sounds interesting is to determine in children, as they grow, if their behavioral product of parental (and otherwise) education are reflected in the shape of their individual scaling function. This seems reasonable given the fact that "tireless running around" is almost a definition of early age well-being, which gives way to less hectic activity as children mature. In the same line of thoughts, if changes in the scaling function can be quantita- tively traced to behavioral changes, one could also consider to explore applications of these techniques to monitor eventual progress in the treatment of hyperactivity disturbances such as in the subjects affected by the Attention Deficit Hyperactivity Dis- order syndrome. The converse, i.e., cases in which the average activity diminish, as in elderly subjects shall be also explored. Further experiments and analysis should shed light on these possibilities. In, the meantime, the present results provide a guide and six important constraints for the models that should best capture the physics (and biology) of the process. Acknowledgements - Work supported by Na- tional Science Center of Poland (ncn.gov.pl, grant DEC-2011/02/A/ST1/00119); State Secretary for Research and Development (grants PSI2010-19372 and PSI2013-48260) from Spain and by CONICET from Argentina.) [1] T Nakamura, K Kiyono, K Yoshiuchi, R Naka- hara, Z R Struzik, Y Yamamoto, Universal scaling law in human behavioral organization, Phys. Rev. Lett. 99, 138103 (2007). [2] T Nakamura, et al., Of mice and men - uni- versality and breakdown of behavioral organi- zation, PLoS ONE 3, e2050 (2008). [3] K Hu, P C Ivanov, Z Chen, M F Hilton, H E Stanley, S A Shea, Non-random fluctuations and multi-scale dynamics regulation of human activity, Physica A 337, 307 (2004). 070017-6 Papers in Physics, vol. 7, art. 070017 (2015) / D. Chialvo et al. [4] L A N Amaral, D J B Soares, L R da Silva, L S Lucena, M Saito, H Kumano, N Aoyagi, Y Ya- mamoto, Power law temporal auto-correlations in day-long records of human physical activ- ity and their alteration with disease, Europhys. Lett. 66, 448 (2004). [5] C Anteneodo, D R Chialvo, Unravelling the fluctuations of animal motor activity, Chaos 19, 033123 (2009). [6] K Christensen, D Papavassiliou, A de Figueiredo, N R Franks, A B Sendova-Franks, Universality in ant behaviour, J. R. Soc. Inter- face 12, 20140985 (2014). [7] A Proekt, J Banavar, A Maritan, D Pfaff, Scale invariance in the dynamics of sponta- neous behavior, Proc Natl Acad Sci USA 109, 10564 (2012). [8] J K Ochab, et al., Scale-free fluctuations in be- havioral performance: Delineating changes in spontaneous behavior of humans with induced sleep deficiency, PLoS ONE 9, e107542 (2014). [9] L Laurson, X Illa, S Santucci, K T Tallak- stad, K J Maloy, M J Alava, Evolution of the average avalanche shape with the universality class, Nature Comm. 4, 2927 (2013). [10] S Papanikolaou, F Bohn, R L Sommer, G Durin, S Zapperi, J P Sethna, Universality beyond power laws and the average avalanche shape, Nature Phys. 7, 316 (2011). [11] J P Sethna, K A Dahmen, C R Myers, Crack- ling noise, Nature 410, 242 (2001). [12] N Friedman, S Ito, B A W Brinkman, M Shi- mono, R E L DeVille, K A Dahmen, J M Beggs, T C Butler, Universal critical dynam- ics in high resolution neuronal avalanche data, Phys. Rev. Lett. 108, 208102 (2012). [13] A Baldassarri, F Colaiori, C Castellano, Av- erage shape of a fluctuation: Universality in excursions of stochastic processes, Phys. Rev. Lett. 90, 060601 (2003). [14] B D Malamud, D L Turcotte, Self-affine time series: I. Generation and analyses, Adv. Geo- phys. 40, 1 (1999). [15] P Bak. How nature works. The science of self- organized criticality, Copernicus, New York (1996). [16] B Gutenberg, C F Richter, Magnitude and en- ergy of earthquakes, Ann. Geofis. 9, 1 (1956). [17] B Wang, J Kuo, S Granick, Burst of active transport in living cells, Phys. Rev. Lett. 111, 208102 (2013). [18] L Laurson, M J Alava, 1/f noise and avalanche scaling in plastic deformation, Phys. Rev. E 74, 066106 (2006). [19] A Clauset, C R Shalizi, M E J. Newman, Power-law distributions in empirical data, SIAM Rev. 51, 661 (2009). [20] F Colaiori, A Baldassarri, C Castellano, Aver- age trajectory of returning walks, Phys. Rev. E 69, 041105 (2004). [21] B B Mandelbrot, Self-affine fractals and frac- tal dimension, Physica Scripta 32, 257 (1985). 070017-7
1812.00278
2
1812
2019-04-26T05:13:15
Interpreting Encoding and Decoding Models
[ "q-bio.NC" ]
Encoding and decoding models are widely used in systems, cognitive, and computational neuroscience to make sense of brain-activity data. However, the interpretation of their results requires care. Decoding models can help reveal whether particular information is present in a brain region in a format the decoder can exploit. Encoding models make comprehensive predictions about representational spaces. In the context of sensory systems, encoding models enable us to test and compare brain-computational models, and thus directly constrain computational theory. Encoding and decoding models typically include fitted linear-model components. Sometimes the weights of the fitted linear combinations are interpreted as reflecting, in an encoding model, the contribution of different sensory features to the representation or, in a decoding model, the contribution of different measured brain responses to a decoded feature. Such interpretations can be problematic when the predictor variables or their noise components are correlated and when priors (or penalties) are used to regularize the fit. Encoding and decoding models are evaluated in terms of their generalization performance. The correct interpretation depends on the level of generalization a model achieves (e.g. to new response measurements for the same stimuli, to new stimuli from the same population, or to stimuli from a different population). Significant decoding or encoding performance of a single model (at whatever level of generality) does not provide strong constraints for theory. Many models must be tested and inferentially compared for analyses to drive theoretical progress.
q-bio.NC
q-bio
Interpreting encoding and decoding models Nikolaus Kriegeskortea & Pamela K. Douglasb a Department of Psychology, Department of Neuroscience, Department of Electrical Engineering, Zuckerman Mind Brain Behavior Institute, Columbia University, New York, NY b Center for Cognitive Neuroscience, University of California, Los Angeles, CA; Modeling, Simulation, Computer Science, UCF, USA Abstract Encoding and decoding models are widely used in systems, cognitive, and computational neuroscience to make sense of brain-activity data. However, the interpretation of their results requires care. Decoding models can help reveal whether particular information is present in a brain region in a format the decoder can exploit. Encoding models make comprehensive predictions about representational spaces. In the context of sensory experiments, where stimuli are experimentally controlled, encoding models enable us to test and compare brain- computational theories. Encoding and decoding models typically include fitted linear-model components. Sometimes the weights of the fitted linear combinations are interpreted as reflecting, in an encoding model, the contribution of different sensory features to the representation or, in a decoding model, the contribution of different measured brain responses to a decoded feature. Such interpretations can be problematic when the predictor variables or their noise components are correlated and when priors (or penalties) are used to regularize the fit. Encoding and decoding models are evaluated in terms of their generalization performance. The correct interpretation depends on the level of generalization a model achieves (e.g. to new response measurements for the same stimuli, to new stimuli from the same population, or to stimuli from a different population). Significant decoding or encoding performance of a single model (at whatever level of generality) does not provide strong constraints for theory. Many models must be tested and inferentially compared for analyses to drive theoretical progress. Highlights 1. Decoding models can reveal whether particular information is present in a brain region in a format the decoder can exploit. 2. Encoding models make comprehensive predictions about representational spaces and more strongly constrain computational theory. 3. The weights of the fitted linear combinations used in encoding and decoding models are not straightforward to interpret. 4. Interpretation of encoding and decoding models critically depends on the level of generalization achieved. 1 5. Many models must be tested and inferentially compared for analyses to drive theoretical progress. Encoding and decoding: concepts with caveats The notions of encoding and decoding are rooted in the view that the brain processes information. Information about the world enters our brains through the senses, is processed and potentially stored, and informs our behavior over a large range of time scales. To understand brain function, then, we must understand the information processing: what information is processed, in what format it is coded in neural activity, and how it is re-coded across stages of processing, so as to ultimately contribute to behavior. The view of the brain as an information-processing device (or, equivalently, a computational device) is closely linked to the concept of representation. A pattern of neural activity represents information about the world if it serves the function to convey this information to downstream regions, which use it to produce successful behavior (DeCharms & Zador 2000). When we talk about encoding and decoding, we focus on a particular brain region X whose function we are trying to understand. To simplify the problem, we divide the process into the encoder (which produces the code in region X from the sensory signals) and the decoder (which uses the code in region X to enable successful behavior). This bipartite division often provides a useful simplification. However, we have to consider three caveats to avoid confusion. First, encoder and decoder do not inherently differ with respect to the nature of the processing. Where the encoder ends and the decoder begins depends on our region of interest X. When we move our focus to the next stage of representation in region Y, the processing between X and Y, which was part of the decoder, becomes part of the encoder. Whether a particular processing step in the brain is to be considered encoding or decoding, thus, is in the eye of the beholder. Second, our region of interest X may not be part of all causal paths from input to output. The division about region X into encoder and decoder, then, misses a portion of the causal graph. Brain regions do not in general form chains of processing stages without skipping connections or recurrent signaling. The primate visual hierarchy is a case in point, where cortical areas interact in a network with about a third of all possible pairwise inter-area connections (Felleman & Van Essen 1991, Young 1992). Encoding and decoding models are nevertheless useful in this context, providing a partial view of the encoded information and its format. Third, the terms encoding and decoding suggest that each is the inverse of the other: The encoder maps stimuli to brain responses; the decoder maps brain responses to stimuli. However, if we would like to interpret our encoding and decoding models as models of brain computation, then both encoder and decoder ought to operate in the causal direction. The decoder should not map back to stimuli, but on to motor representations. An encoding model might predict neural activity elicited by images (Wu et al. 2006, Kay et al. 2008, Wen et al. 2018). A decoding model might attempt to predict category labels from neural activity, thus explicating information that might be reflected in a behavioral response (Majaj et al. 2015). The notion that encoding is followed by its inverse begs the question why the original input is not simply copied. A possible motivation for a sequence of an encoder and its inverse is to temporarily compress the information, e.g., for passage through a bottleneck such as the optic nerve. A more general notion is that information is re-coded through a sequence of stages, possibly compressing it for more efficient representation (Simoncelli & Olshausen 2001), but 2 also discarding unneeded information and converting needed information into a format that enables it to be exploited to control behavior (Tishby et al. 2000, Achille & Soatto 2018). When decoding models are conceptualized as inverse encoders (mapping back to the stimuli, rather than on to downstream representations), they cannot be interpreted as process models of brain function. However, they can still be useful tools, revealing information present in a brain region and giving us clues about the format of the code. For a careful discussion of the interpretation of causal and anti-causal encoding and decoding models, see Weichwald et al. (2015). Encoding and decoding are important concepts in both theoretical and experimental neuroscience. This paper addresses experimentalists and focuses on the interpretation of empirical results obtained by fitting encoding and decoding models to brain-activity data. Decoding models: revealing information and its format Decoding is sometimes described as "reading the brain" or "cracking the neural code" (Cox & Savoy 2003, Haynes & Rees 2006, Tong & Pratte 2012). The underlying concept is that of decoding as inverse encoding, where the goal is not to model brain information processing, but to reveal the content of the code. The sense of revealing a mystery and gaining insight into hidden information drives the imagination of scientists and lay people alike. The intuition that something important has been learned when we can decode some perceptual or higher cognitive content from brain activity is correct. However, "decrypting" the brain should not be equated with understanding its computational mechanism. Decoding reveals the products, not the process of brain computation (Box 1). However, it is a useful tool for testing whether a brain region contains a particular kind of information in a particular format (Haxby et al. 2001, Cox & Savoy 2003, Carlson et al. 2003, Kamitani & Tong 2005, Haynes & Rees 2006, Kriegeskorte et al. 2006, Norman et al. 2006, Paninski et al. 2007, Mur et al. 2009, Pereira et al. 2009, Pillow et al. 2011, Kriegeskorte 2011, Tong & Pratte 2012, Hebart et al. 2015, Haynes 2015, Hebart & Baker 2017, Varoquaux et al. 2017, Paninski & Cunningham 2017). A simple example of a decoding model is a linear classifier (Duda et al. 2012) that takes a measured brain-activity pattern as input and outputs a class label (say "cat" and "dog"), revealing which of two stimuli (say a particular image of a cat and a particular image of a dog) elicited the response pattern (Figure 1). A linear classifier may achieve this by computing a weighted sum of the measured responses and imposing a threshold to decide which label to return (see Mur et al. 2009 for an introduction in the context of neuroimaging). The weights are set to maximize the accuracy of the decoding on a training data set. If decoding succeeds reliably on a test data set, then the region must contain information about the decoded variable (the binary stimulus distinction here). Decoding provides a statistically advantageous way of testing for stimulus information In the two-stimulus example, all the linear decoder shows is that the two images elicit distinct response patterns. This means that there is mutual information between stimulus and response. To demonstrate mutual information, we could have used an encoding model (Naselaris et al. 2011) or a multivariate test of difference between the response patterns elicited by the two 3 stimuli (e.g. multivariate analysis of variance), instead of a decoding model. However, a univariate encoding model would in general have less sensitivity, because it does not account for the noise correlations between different response channels (Averbeck et al. 2006). A multivariate analysis of variance would account for the noise correlations, but might have less specificity. In fact, it might fail to control false-positives at the nominal level if its assumptions of multivariate normality did not hold (as is often the case), making it invalid as a statistical test (Kriegeskorte 2011). Decoding provides a natural approach to modelling the noise correlations (e.g. using a multivariate normal model as in the Fisher linear discriminant), without relying on the model assumptions for the validity of the test: Violations of the decoding model assumptions will hurt decoding performance. We, thus, err on the safe side of concluding that there is no information about the stimuli in the responses. In sum, it is not the direction of the decoding model ("brain reading") that makes it a compelling test for information, but the statistical nature of the problem (noise correlations) and the fact that decoders are tested on independent data. Figure 1 Linear encoding and decoding models. (A) Encoding and decoding model the relationship between stimuli and responses in opposite directions. An encoding model (top) predicts brain responses as a linear combination of stimulus properties (black circles). A decoding model (bottom) predicts stimulus properties as a linear combination of brain responses. (B) Example of a linear decoding model using a 2- dimensional feature space consisting of two voxels. Voxel 1 contains relevant information about which of two classes (green, blue) the stimulus belongs to. Voxel 2 contains no information about the stimulus class. The two dimensions jointly define the linear discriminatory boundary. Note that the weights assigned to each voxel are defined by the vector w, which is orthogonal to the decision boundary. Because the noise is correlated between the voxels, a linear decoder will assign significant negative weight to voxel 2, using this voxel (which contains only noise) to cancel the noise in the voxel which contains signal. As a result, interpreting the absolute weights of linear decoders requires care and additional analyses. Linear decodability indicates "explicit" information For decoding to succeed, the information must be present in the brain region in a format that the decoder can exploit. Linear decoders, the most widely used class, require that the distributions of patterns be linearly separable to some extent. This is a weakness in that we might fail to detect information encoded in a more complex format. However, it is a strength in that it provides clues to the format of the information we do detect. The simpler the decoder, the more focused its sensitivity will be. From the perspective of understanding the brain computations, it 4 is attractive to use decoding operations that single neurons can implement. These include linear readout, but also simple nonlinear forms of readout such as radial basis function decoding (Poggio & Girosi 1990). Linear decodability indicates that a downstream neuron receiving input from a sufficient portion of the pattern, could read out the information in question (DiCarlo & Cox 2007). Information amenable to direct readout by downstream neurons is sometimes referred to as "explicit" in the code (Kriegeskorte 2011, DiCarlo et al. 2012, Hong et al. 2016). The level of generalization beyond the training set must be considered when interpreting a decoding result Fitting a model always poses the risk of overfitting, i.e. of optimizing the fit to the training data at the expense of predictive performance on independent data. Overfitting can lead to high decoding accuracy on the training set, even if the response patterns contain no information about the stimulus. Decoders therefore need to be tested for generalization to independent data (Hastie et al. 2009, Kriegeskorte 2015). In our example, we might test the decoder on an independent set of response measurements for the same two particular images of a cat and a dog. If decoding accuracy on this independent test set is significant, we can reject the null hypothesis that the response patterns contain no information about the stimulus (Mur et al. 2009, Pereira et al. 2009). However, detecting information about which of two images has been presented tells us almost nothing about the nature of the code. The two images must have distinct response patterns in the retina and V1 (low-level representations) as well as in the visuo-semantic regions of the ventral stream (high-level representations). We would therefore expect a linear decoder to work on new measurements for the same images in any of these regions. This reflects the fact that all the regions contain image information. In the retina, for example, we expect the two images to elicit distinct response patterns, while the manifolds of response patterns corresponding to the two categories are hopelessly entangled (DiCarlo & Cox 2007, Chung et al. 2016, Chung et al. 2018). Given responses to just two images, we can demonstrate the presence of information, but have no empirical basis for characterizing the nature of the code (see Kriegeskorte et al. 2007 for a study limited by this drawback). In order to learn whether the region we are decoding from contains a low-level image encoding or a high-level categorical encoding, we can train the decoder on one set of cat and dog images and test it on another set of images of different cats and dogs (Anzellotti et al. 2014, Freedman et al. 2002, Kriegeskorte 2011). To support the interpretation that "cats" and "dogs" are linearly separable in the representation (rather than the weaker claim that there is image information), it is not sufficient to increase the number of particular images of cats and dogs, while training and testing on the same images. The linear decoder has many parameters (one for each response channel) and is expected to overfit even to a larger set of particular images. Even for the retinal representation, we therefore expect a cat/dog decoder to generalize to new measurements performed on the same images. We must test the linear decoder for generalization to different cats and dogs. Note, however, that interpreting linear decodability as linear separability of the two classes in the neuronal representational space would further require the decoding accuracy to be so high that errors can be attributed to the measurement noise rather than the neural representation. In practice, we typically face ambiguity. For example, decoding accuracy may be significantly 5 above chance, but far from perfect. This indicates that the code contains some linearly decodable information, but claims of linear separability may be difficult to evaluate as it would require attributing the substantial proportion of errors to limitations of the measurements (noise and subsampling), rather than to a lack of linear separability of the neuronal activity patterns. If we managed to show that cats and dogs fall in linearly separable regions in representational space, there would still be a question about the nature of the features that support the encoding. There may be many different visual features that can be computed from images and that lend themselves to separating cats from dogs (e.g. the shape of the ears or the shape of the eyes). Revealing which particular features the encoding in a given brain region employs would require further experiments. For example, we could design artificial stimuli that contain subsets of the distinctive features of cats and dogs. We could then test whether a linear decoder trained on responses to natural images of cats and dogs generalizes to responses to the artificial stimuli. Successful decoding would support the hypothesis that the brain region employs some of the features present in the artificial images. More generally, we can go beyond using different exemplars from the same categories in the test set. Testing decoders for cross-generalization between different domains is useful to address a host of questions about the invariances of the encoding, for example its consistency between imagery and perception (Stokes et al. 2009, Cichy et al. 2012) and its stability across latencies after stimulus onset (King & Dehaene 2014). Stimulus reconstruction provides a richer view of the information present, but complex priors complicate the interpretation A decoder may discriminate more than two categories. It might decode a continuous variable (e.g. the orientation of the stimulus; Kamitani & Tong 2005), or it might output a multidimensional description of the stimulus. The most ambitious variant of decoding is stimulus reconstruction, where the output is a detailed estimated rendering of the stimulus. A decoder providing a rendering of the stimulus can potentially reveal more of the information present in the neuronal encoding (Thirion et al. 2006, Miyawaki et al. 2008, Mesgarani et al. 2009, Naselaris et al. 2009, Santoro et al. 2017, Parthasarathy et al. 2017). Stimulus reconstruction is a particularly impressive feat of decoding for two related reasons. First, the space of outputs is much richer, so the decoder reveals more of the encoded information (discriminating among more possible stimuli). Second, a more challenging level of generalization is typically required for the decoder to discriminate among this richer set of possible stimuli, because there are usually severe limitations on the number of stimuli that the decoder can be directly trained on. Successful generalization suggests that the structure of the model captures something about the code that enables generalization far beyond the training set. To the extent that stimulus reconstruction, when applied to novel stimuli that the model has not been trained on, successfully specifies close matches within a rich space of reconstruction targets (e.g. pixel images), the analysis indicates that the code contains rich information about the stimulus. An extreme example of this would be a decoder that can precisely reconstruct arbitrary natural or unnatural images from their representation in V1. This would indicate that V1 encodes the image precisely and that the encoding is not restricted to natural images. In fact, 6 this has never been shown for V1 and would be puzzling, because we expect visual representations to be specialized for natural stimuli. Box 1: Decoding Models: Benefits and Limitations Decoding provides an intuitive and compelling demonstration of the presence of information in a brain region. Decoding models bring several benefits: • They enable researchers to look into the regions and reveal whether particular information is encoded in the responses. Encoded information might be used by downstream neurons (i.e. the encoding might serve as a representation). Training and testing decoders has helped move the focus from single-neuron activity and regional-mean activation to the information being processed, making analyses more relevant to the computational function in question. • They enable cognitive neuroscience to exploit the fine-grained multivariate information present in modern imaging data. Before the advent of decoding, dominant brain mapping techniques required smoothing of the data, treating fine-grained information as noise. The multivariate modelling of local patterns also brought locally multivariate noise models to the literature, which greatly enhance the information that can be drawn from neuroimaging data. • Decoding analyses use independent test data to assess the presence of information. Demonstrating significant information with independent test data ensures that violations of model assumptions lead to conservative errors (making significant results less likely). This is an advantage over methods (such as univariate and multivariate linear regression) that use distributional assumptions rather than independent test sets to perform statistical inference. Decoding models are limited in several ways: • They cannot in general be interpreted as models of brain-information processing. In the context of sensory systems, decoding models operate in the wrong direction, capturing the inverse of the brain's processing of the stimuli. If the decoded variable is a behavioral response, then direction of information flow matches between model and brain. However, such applications are rare. Moreover, the most successful applications of decoding in the literature rely on linear decoders, which are useful to reveal explicit information, but cannot capture the complex nonlinear computations we wish to understand. • Decoder weight maps are difficult to interpret (Figure 3) because voxels uninformative by themselves can receive large weights when they help cancel noise and because the weights are codetermined by the data and the prior, and the fitting procedure might select one but not the other of two essentially equally informative voxels. • Decoding results provide only weak constraints on computational mechanisms. Knowing that a particular kind of information is present provides some evidence for and against alternative theories about what the region does. However, it does not exhaustively characterize the representation, let alone reveal the underlying computations. In order to improve the appearance of reconstructions of natural stimuli, it is attractive to use a prior over the possible outputs of the decoder. For example, we might constrain the model to output an image whose low-level or high-level statistics match natural images. We may constrain the decoder more strongly, by limiting it to output only actual natural images. This constraint could be implemented using a restricted set of target images (Naselaris et al. 2009) or a generative model of natural images (Goodfellow et al. 2014). In either case, the quality of the reconstruction will reflect a combination of the information provided by the brain region and the information contributed by the prior. A complicated prior therefore also complicates the 7 interpretation of the reconstruction results: good looking reconstructions no longer indicate that all the detail they provide is encoded in the brain region. The reconstruction has to be compared to the presented stimulus, and the complexity of the output space (which is reduced by the prior over the outputs) needs to be considered in the interpretation. An important question is what we can learn from stimulus reconstructions. The goal to learn about the content and format of the code may not be ideally served by striving for the most natural looking reconstruction. Decoding models predicting behavioral responses from brain responses can be interpreted as brain-computational models Decoding is usually used as a tool of analysis that reveals aspects of the content and format of the information encoded in a brain region. The decoding model, thus, is not interpreted as a model of brain computation. In the context of sensory systems, a decoder maps from brain responses to stimuli. Since stimulus processing by the brain operates in the opposite direction, it is difficult to interpret a decoder as a model of brain information processing. However, if a decoder is used to predict behavioral responses, e.g. judgments of categorical or continuous stimulus variables (possibly including errors and reaction times on individual trials), then the decoder can be interpreted as a model (at a high level of abstraction) of the brain computations generating the behavioral responses from the encoding of the stimuli in the decoded brain region (Shadlen et al. 1996, Williams et al. 2007, Walther et al. 2012). Encoding models: testing comprehensive representational predictions Encoding models attempt to predict brain response patterns from descriptions of the experimental conditions (Figure 1A; Paninski et al. 2007, Kay et al. 2008, Dumoulin & Wandell 2008, Mitchell et al. 2008, Naselaris et al. 2011, Khaligh-Razavi & Kriegeskorte 2014, Yamins et al. 2014, Naselaris & Kay 2015, van Gerven 2017). Encoding models, thus, operate in the opposite direction as decoding models. If our goal is merely to demonstrate that a brain region contains information about the experimental conditions, then the direction the model should operate in is a technical issue: One direction may be more convenient for capturing the relevant statistical dependencies (e.g. noise correlations among responses), but a model operating in either direction could support the inference that particular information is present in the code. If our goal is to test computational theories, however, then the direction that the model operates in matters, because it determines whether the model can be interpreted as a brain-computational model. Encoding models predicting brain responses from sensory stimuli can serve as brain-computational models Whereas a decoding model typically serves to test for the presence of particular information in a brain region, an encoding model can provide a process model, at some level of abstraction, of the brain computations that produce the neuronal code. An encoding model makes 8 comprehensive predictions about the representational space (Naselaris & Kay 2015, Diedrichsen & Kriegeskorte 2017; Box 2) and, ideally, should fully explain neuronal responses in the region in question (up to the maximum achievable given the noise in the data). In sensory neuroscience, the experimental conditions correspond to sensory stimuli, and so an encoding model maps sensory stimuli to their encodings in sensory regions of the brain (Kay et al. 2008, Mitchell et al. 2008, Dumoulin & Wandell 2008, Huth et al. 2012). A sensory encoding model, thus, operates in the direction of the information flow: from stimuli to brain responses. It should take raw sensory stimuli as input (e.g. images or sounds; Kay et al. 2008, Naselaris et al. 2011, Khaligh-Razavi & Kriegeskorte 2014, Yamins et al. 2014, Kell et al. 2018; for recent reviews see Kriegeskorte 2015, Yamins & DiCarlo 2016) and predict the patterns of brain activity the stimuli elicit. We can adjudicate among computational theories by implementing them in encoding models and testing their ability to predict brain-activity patterns (Kriegeskorte & Diedrichsen 2016). Brain-computational encoding models can be tested by predicting raw measurements, representational dissimilarities, or the activity-profiles distribution A brain-computational encoding model could consist in a set of hand-engineered features computed from the stimuli or in a neural network model trained on some task. How can we evaluate a high-dimensional representation in a brain-computational encoding model with brain- activity data, when the units of the model may not correspond one-to-one to the measured channels of brain activity? One approach is to predict the raw measurements (Kay et al. 2008, Dumoulin & Wandell et al. 2008, Mitchell et al. 2008, Naselaris et al 2011, van Gerven 2017). To achieve this, we can fit a linear model that maps from the outputs of the nonlinear encoder thought to capture the brain computations (e.g. a layer of a neural network model processing the stimuli) to each measured response channel, e.g. each neuron or fMRI voxel in the region we would like to understand. The linear model will capture which units of the nonlinear encoding model best predict each response channel. It can be interpreted as capturing the sampling and averaging that occurs in the measurement of brain activity. A neuronal recording array, for example, will capture a sample of neurons, and fMRI will give us measurements akin to local spatiotemporal averages at regular grid locations. For each response channel, the linear model will have a parameter for each nonlinear feature (e.g. each unit of the neural network layer). Fitting these parameters requires substantial training data and normally also a prior on the parameters (e.g. a 0-mean Gaussian prior as is implicit to fitting with an L2 penalty). Instead of predicting the raw measurements, we can use a brain-computational encoding model to predict to what extent different stimuli are dissimilar in the representation, an approach called representational similarity analysis (RSA; Kriegeskorte et al. 2008, Nili et al. 2014). The pairwise dissimilarities of the multivariate response patterns representing the stimuli can be summarized in a representational dissimilarity matrix (RDM). For example, for each pair of stimuli, the dissimilarity between the associated response patterns could be measured using the Euclidean distance. The RDM characterizes the geometry of the set of points in the multivariate response space that correspond to the stimuli. Noise correlations can be captured by estimating the covariance matrix of the residuals of the response-estimation model, and replacing the Euclidean distance by the Mahalanobis distance. To remove the positive bias associated with 9 measuring distances between noisy data points, we can use the crossnobis (crossvalidated Mahalanobis) estimator (Nili et al. 2014, Walther et al. 2016). The resulting crossnobis RDM provides a full characterization of the linearly decodable information in the representational space (Diedrichsen & Kriegeskorte 2017). Comparing representations in models and brains at the level of RDMs obviates the need for fitting a linear model to predict each measured response (thus reducing the need for training data) and enables the analysis to naturally handle noise correlations between responses (which are typically ignored when encoding models separately predict each of the measured response channels). A third approach to the evaluation of encoding models is to predict the distribution of activity profiles. In pattern component modelling (PCM, Diedrichsen et al. 2011), this distribution is characterized by the second moment of the activity profiles. Like the RDM, this is a stimulus-by- stimulus summary statistic of the stimulus-response matrix. Each entry of the second-moment matrix corresponds to the inner product between two response patterns. All three approaches can be construed as testing hypotheses about the representational space induced by the activity profiles (Diedrichsen & Kriegeskorte 2017). Consider a linear encoding model using a Gaussian prior on the weights. Such a model predicts a Gaussian distribution of activity profiles. The predicted distribution of activity profiles is captured by its second moment. For representational similarity analysis, similarly, the RDM is a function of the second moment of the activity profiles. This core mathematical commonality between the methods notwithstanding, each is best suited for a particular set of questions. Linear models predicting raw measurements lend themselves to univariate brain mapping, revealing which voxels or neurons are accounted for by a particular nonlinear encoding model. RSA lends itself to characterizing the geometry of the representational space, naturally handles noise correlations among responses, and reduces the need for training data. PCM can have greater sensitivity for adjudicating among models than the other two methods, at the expense of relying on stronger assumptions. The three methods are best viewed as part of a single toolbox of representational model analyses, whose elements can be combined as needed to address particular questions. The level of generalization beyond the training set must be considered when interpreting an encoding result Encoders, like decoders, are tested by evaluating how well they predict independent data, whether the predicted quantities are the raw brain-activity measurements, the representational dissimilarity matrix, or the second-moment matrix of the activity profiles. For encoders, as for decoders, the interpretation depends on both the prediction accuracy and the level of generalization beyond the training set that the model achieves. Encoding models typically require the fitting of parameters, so overfitting needs to be accounted for in any inferential procedure. In the simplest type of a univariate linear encoding model, we can rely on Gaussian assumptions and perform inference without a separate test set (e.g. Friston et al. 1994). However, more interesting models require independent test sets, for example when parameters are fitted using priors over the weights and when the model is a brain-computational model to be tested for generalization to new conditions. 10 A key consideration is how much flexibility to allow in fitting each model representation to a brain representation. One extreme is to allow no flexibility and assume that the model representation precisely predicts the geometry of the representational space (Kriegeskorte et al. 2008). This case is most naturally handled by RSA and PCM, but could also be implemented with linear encoders by using a prior that prevents any distortion of the representational geometry. The other extreme is to allow arbitrary linear remixing of the units of the nonlinear encoding model. This case is most naturally handled with linear encoding models, but can also be implemented with PCM and RSA (Diedrichsen et al. 2011, Khaligh-Razavi & Kriegeskorte 2014, Khaligh-Razavi et al. 2016). In practice, some compromise is desirable, which we can think of as a prior on the mapping from the brain-computational model to the measured brain responses. We might use a 0-mean Gaussian prior on the weights (e.g. Kay et al. 2008). Alternatively, we can limit flexibility more aggressively, by allowing each unit (or each feature map or layer) a single weight (not a separate weight for each response). Such weighted representational models (e.g. Khaligh-Razavi et al. 2014) are naturally implemented with RSA and PCM. Each brain-computational model in this case predicts a superset of the features spanning the brain representational space (disallowing linear mixtures), but does not predict the prevalence of each of the features in the neuronal population. The lowest level of generalization beyond the training set is generalization to new measurements for the same experimental conditions. This is sufficient, if the experimental conditions exhaustively cover the domain we would like to draw inferences about (consider the case of the representation of the five fingers in motor cortex: Diedrichsen et al. 2012). However, in a domain such a sensory systems, the goal is typically to evaluate to what extent a brain- computational model can predict brain representations of arbitrary stimuli. This requires a higher level of generalization beyond the training set. A vision model, for example, might be trained with responses to one sample of natural images and tested for generalization to responses to an independent (and nonoverlapping) sample from the same distribution of natural images. Because the set of all natural images is so rich, this is a challenging generalization task (as illustrated by the difficulty of computer vision). An even more stringent test of the assumptions implicit to a model is to train the model on a sample from one population of images and test it on a sample from a different population of images (e.g. Eickenberg et al. 2017). The prediction accuracy can be assessed in terms of whether it is significantly above chance level, whether it significantly differs from that of competing encoding models, and how close it comes come to the noise ceiling (the highest achievable accuracy given the noise in the data, Nili et al. 2014). We can generalize claims about an encoding model to the extent that its predictions generalize. If we want to conclude that the model can predict responses for the stimuli presented, we need not test the model with different stimuli (only with different response measurements elicited by the same stimuli). If we want to conclude that the model can predict responses to arbitrary natural stimuli, we need to test it with new arbitrary natural stimuli. The population of conditions the test set is a sample of defines the scope of the claims we can make (Hastie et al. 2009, Kriegeskorte 2015). We focus here on encoding and decoding models that are fitted to individual subjects' brains, so as be able to exploit fine-grained idiosyncratic patterns of activity that are unique to each subject. Within-subject prediction accuracy may support generalization to a population of stimuli, but it doesn't support generalization to the population of subjects. In some fields, such as low- 11 level vision, researchers draw on prior knowledge and assume that results that hold for a few subjects will hold for the population. If we were instead to use our data to generalize our inferences to the population of subjects, we would either need to predict results for held-out subjects (which would make it impossible to exploit individually unique fine-grained activity patterns) or perform inference on the within-subject prediction accuracies with subject as a random effect. The feature fallacy: Interpreting the success of a model as confirmation of its basis set of features Linear encoding models predict each measured activity profile as a linear combination of a set of model features. When a model can explain the measured activity profiles, we might be tempted to conclude that the model features are encoded in the measured responses. This interpretation is problematic because the same linear space can be spanned by many alternative basis sets of model features. The fact that multiple sets of basis vectors can span the same space is widely appreciated. However, it is not obvious whether the ambiguity is removed when a prior over the encoding weights is used. An encoding model with a 0-mean isotropic Gaussian prior on the weights (equivalent to ridge regression) predicts a Gaussian distribution of activity profiles (captured by the second moment of the activity profiles as the sufficient statistic) and a particular representational geometry (captured by the RDM). The use of a weight prior does reduce the ambiguity. In the absence of a weight prior, all models spanning the same space of activity profiles make equivalent predictions. With a weight prior, different models spanning the same space can predict distinct distributions of activity profiles. However, substantial ambiguity remains (Figure 2) because there are infinitely many alternative ways to express the same distribution of activity profiles by different feature sets (each assuming a 0-mean isotropic Gaussian prior over the weights). The freedom to choose among feature sets generating the same distribution of activity profiles is useful in that it enables us to implement a nonisotropic Gaussian prior on the weights of a given model (Tikhonov regularization) by re-expressing the model such that the features induce the same distribution of activity profiles in combination with a 0-mean isotropic Gaussian prior (Nunez-Elizalde et al. 2018). However, this freedom to express the same model by different feature sets needs to be kept in mind when interpreting results: Many alternative feature sets would have given the same results. In sum, the same distribution of activity profiles can be expressed in many ways (by different feature sets), so the fact that a model explains brain responses does not provide evidence in favor of the particular feature set chosen. Diedrichsen (2018) termed the interpretation in terms of the particular features used the "feature fallacy". This fallacy is arguably somewhat persistent in the neuroscience literature (Churchland et al. 2012). 12 Figure 2 The feature fallacy. Different linear encoding models spanning the same space of activity profiles may not be distinguishable. There are many alternative sets of feature vectors {f1, f2, …} that span the same space of activity profiles. In the absence of a prior on the weights of the linear model, all these sets can equally explain a given set of brain responses. The ambiguity is reduced, but not resolved when a prior on the weights is assumed (Diedrichsen et al. 2018). If we define a prior on the weights, then each model predicts a probability density over the space of activity profiles. This probabilistic prediction may be distinct for two sets of basis features, even if they span the same space. For example, if the weight prior is a 0-mean isotropic Gaussian, then each model assigns probabilities to different activity profiles according to a Gaussian distribution over the space of profiles. Two linear models may span the same space, but predict distinct distributions of activity profiles. However, even with a Gaussian weight prior, there are still (infinitely) many equivalent models that make identical probabilistic predictions. We illustrate this by example. (a) Three models (A, B, C) each contain two feature vectors as predictors (A: {fA1, fA2}, B: {fB1, fB2}, C: {fC1, fC2}). The three models all span the same 2-dimensional space of activity profiles. For each model, we assume a 0-mean isotropic Gaussian weight prior. (b) All three models predict the same nonisotropic Gaussian probability density over the space of activity profiles (indicated by a single iso-probability-density contour: the ellipse). Model A (gray) predicts the density by modeling it with two orthogonal features that capture the principal-component axes, with features having different norms to capture the anisotropy. Model B predicts the same density by modeling with two correlated features of similar norm. Model C falls somewhere in between, combining feature correlation and different feature norms to capture the same Gaussian density over the activity profiles. Note that there are many other models that span the same space, but will not induce the same probability density over activity profiles when complemented by a 0- mean isotropic Gaussian weight prior. A given linear encoding model's success at predicting brain responses provides evidence for the induced distribution of activity profiles, but not for the particular features chosen to express that distribution. 13 Box 2: Encoding Models: Benefits and Limitations An encoding model predicts the response of each measurement channel (e.g. a neuron or fMRI voxel) on the basis of properties of the experimental condition (e.g. sensory stimulus). Continuous brain maps can be obtained by fitting such a model to each response in turn. Responses are typically predicted as linear combinations of the features, rendering this approach closely related to classical univariate brain mapping. However, encoding models have several important benefits: • More complex feature sets are used, often comprising thousands of descriptive features. Fitting the models therefore requires regularization. Typically, models are fitted using a penalty on the weights. • Encoding models can have nonlinear components, such as the location and size of a visual receptive field. • Models are tested for generalization to different experimental conditions (e.g. a different sample of visual stimuli). • When sensory systems are studied, encoding models operate in the direction of information flow. They can then be used as a general method for testing brain-computational models, i.e. models that process sensory stimuli. A vision model, for example, can take a novel image as input and predict responses to that image in cortex. Analyses using encoding models to predict each response channel separately are limited in two ways: • The separate modeling of each response channel is inconsistent with the multivariate nature of neuronal population codes and noise, and is statistically suboptimal. Separate modeling of each response entails low power for testing and comparing models, for two reasons: (1) Single responses (e.g. fMRI voxels) may be noisy and the evidence is not combined across locations. (2) The analyses treat the responses as independent, thus forgoing the benefit exploited by linear decoding approaches to model the noise multivariately. This is important in fMRI, where nearby voxels have highly correlated noise. As spatial resolution increases, we face the combined challenge of more and noisier voxels. This renders mapping with proper correction for multiple testing very difficult. In addition, relating results between individuals is not straightforward. • When model parameters are color-coded and mapped across the cortex, the resulting detailed maps are not straightforward to interpret. (1) Weight maps of linear models are problematic to interpret in encoding models for the same reason they are problematic to interpret for decoding models: because a predictor's weight depends on the other predictors present in the model (unless each predictor is orthogonal to all others). The required regularization further complicates weight interpretation. (2) Models are fit at many locations and voxels highlighted on the basis of model fits. Post-selection inference on parameters is not usually performed. Because of these complications, results of fine-grained mapping across voxels are difficult to substantiate by formal inference. Influential studies have met these challenges by focusing on single-subject analyses, acquiring a large amount of data in each subject, and limiting formal inference with correction for multiple testing to the overall explained variance of a model, with colors serving exploratory purposes. 14 Weights of linear models are not straightforward to interpret in either encoding or decoding models Beyond interpretation of the overall success of an encoding or decoding model, researchers often want to dig deeper and interpret the fitted parameters of their models. In the context of a decoding model, the weights assigned to the voxels would seem to tell us where the information the decoder uses is located in the brain. Similarly, in the context of an encoding model, the weights of different model features promise to reveal to what extent different model features are encoded in a brain region. Unfortunately, the interpretation of the weights of linear models is not as straightforward as the simplicity of a linear relationship might suggest. A weight does not reflect the predictive power of an individual predictor (a measured brain response in decoding, a model features in encoding). Rather a weight reflects a predictor's contribution in the context of the rest of the model. Uninformative predictors can receive large positive or negative weights. For example, an fMRI voxel that does not by itself contain any information can have a large absolute weight in a linear decoder if it improves decoding accuracy by cancelling noise that the voxel shares with other voxels that do contain stimulus information (Figure 1B; Figure 3; Haufe et al. 2014). In an encoding model, similarly, a model feature might contain no information about the modeled response and still receive a large absolute weight. For example, an fMRI voxel that only responds to faces might be explained by a set of semantic descriptors unrelated to faces if the nonface semantic descriptors are capable, in combination, of capturing contextual variation that is correlated with the presence of faces. Conversely, informative predictors may receive weights that are small in absolute value or zero. For example, in fMRI, a voxel might receive zero weight when other voxels suffice for decoding when a weight penalty (especially a sparsity encouraging weight penalty, e.g. defined by the L1 norm of the weights) leads the fitting procedure to select among equally informative voxels (Figure 3). Weight penalties can similarly suppress model feature weights in encoding models. A related problem is performing statistical inference to test hypotheses about the weights (Taylor & Tibshirani 2015). Systems and cognitive neuroscience has yet to develop a proper set of methods for hypothesis testing in this context. A simple remedy to the complications associated with the interpretation of fitted weights is to interpret only the accuracy of decoding and encoding models (and its significance level) and not the parameters of the models. In the context of decoding, this makes sense for local regions of interest corresponding, for example, to cortical areas, which we can test one by one for the presence of particular information. It can be generalized to brain mapping by applying the decoder (or more generally any multivariate analysis) within a searchlight that scans the brain for the effect of interest (Kriegeskorte et al. 2006). Like classical univariate brain mapping, this approach derives its interpretability from the fact that each location is independently subjected to the same analysis. However, instead of averaging local responses, the evidence is summarized using local multivariate statistics. In the context of encoding models, similarly, we can focus on each model's overall performance and on inferential comparisons among multiple models. 15 Figure 3 Weights of decoding and encoding models are difficult to interpret. Three examples (rows) illustrate the difficulty of interpreting decoder weights for a pair of voxels. In the first example (top row), only the top voxel contains signal (stimulus information, red) and the two voxels have independent noise. This scenario is unproblematic: both univariate mapping (second column from the left) and decoder weight maps detect the informative voxel (red). In the second example (second row), both voxels contain the same signal. Here, univariate mapping and weight maps often work. However, the LASSO decoder, because of its preference for a sparse solution, may choose one of the voxels arbitrarily. In the third example (third row), only the top voxel contains signal and both voxels contain correlated noise. Univariate mapping correctly identifies the informative voxel. Linear decoders will give negative weight (blue) to the uninformative voxel, so as to cancel the noise. The single-model-significance fallacy When our goal is merely to detect information in a brain region, we don't interpret the model as a model of brain computation. This lowers the requirements for the model: It need not operate in the direction of information flow and it need not be neurobiologically plausible. The model is merely a statistical tool to sense a dependency. The choice of model in this scenario, will affect our sensitivity, and its structure is not entirely irrelevant to the interpretation. For example, decodability by a linear model tells us something about the format of the encoding. However, a single model will suffice to demonstrate the presence of information in a brain region. When our goal is to gain insight into the computations the brain performs, a model has a more prominent role: it is meant to capture, at some level of abstraction, the computations occurring in the brain. We then require the model to operate in the causal direction and to be neurobiologically plausible (albeit abstracted). Examples of such models include encoding models of sensory responses and decoding models that predict behavioral responses from brain activity. Psychophysical models, which skip the brain entirely and predict behavioral responses directly from stimuli, also fall into this class. In all these cases, finding that a model explains significant variance is a very low bar and tells us little as to whether the model captures the computational process. 16 The single-model-significance fallacy is to interpret the fact that a single model explains significant variance as evidence in favor of the model. A simple example that is widely understood is linear correlation. A significant linear correlation does demonstrate a dependency between two variables, but it does not demonstrate that the dependency is linear. Similarly, the fact that a complex encoding model explains significant variance in the responses of a brain region to a test set of novel stimuli does demonstrate that the brain region contains information about the stimuli, but it does not demonstrate that the encoding model captures the process that computes the encoding. Even a bad model can explain significant variance, especially if it has a large number of parameters fitted to the data. In order to learn about the underlying brain computations, we need to (a) consider multiple models, (b) assess what proportion of the explainable (i.e. nonnoise) variance each explains at a given level of generalization, and (c) compare the models inferentially. Representational interpretations require additional assumptions Decoding and encoding models are often motivated by the goal to understand how the brain represents the world, as well as the animal's decisions, goals, plans, actions, and motor dynamics. Significant variance explained by encoding and decoding models demonstrates the presence of information. Interpreting this information as a representation (Dennett 1987) implies the additional claim that the brain activity serves the purpose to convey the information to other parts of the brain (Kriegeskorte & Bandettini 2007, Kriegeskorte 2011, Diedrichsen & Kriegeskorte 2017). This functional interpretation is so compelling in the context of sensory systems that we sometimes jump too easily from findings of information to representational interpretations (de Wit et al. 2016, Ritchie et al. 2017). In addition to the presence of the information, its functional role as a representation implies that the information is read out by other regions, affecting downstream processing and ultimately behavior. Combining encoding and decoding models with stimulus- and response-based experimentation can help disambiguate the causal implications (Weichwald et al. 2015). Ideally, experimental control of neural activity should also be used to test whether activity has particular downstream or behavioral consequences (Afraz et al. 2006, Raizada & Kriegeskorte 2010). To the extent that we rely on prior assumptions to justify a representational interpretation, it is important to reflect on these and consider if there is evidence from previous studies to support them. 17 References 1. Achille, A., & Soatto, S. (2018). Information dropout: Learning optimal representations through noisy computation. IEEE transactions on pattern analysis and machine intelligence, 40(12), 2897-2905. 2. Afraz, S. R., Kiani, R., & Esteky, H. (2006). Microstimulation of inferotemporal cortex influences face categorization. Nature, 442(7103), 692. 3. Anzellotti, S., Fairhall, S. L., & Caramazza, A. (2014). Decoding Representations of Face Identity That are Tolerant to Rotation. Cerebral Cortex, 24(8), 1988 -- 1995. https://doi.org/10.1093/cercor/bht046 4. Averbeck, B. B., Latham, P. E., & Pouget, A. (2006). Neural correlations, population coding and computation. Nature reviews neuroscience, 7(5), 358. 5. Carlson, T. A. et al. (2003) Patterns of activity in the categorical representation of objects. J. Cogn. Neurosci. 15, 704 -- 717 6. Churchland, M. M., Cunningham, J. P., Kaufman, M. T., Foster, J. D., Nuyujukian, P., Ryu, S. I., & Shenoy, K. V. (2012). Neural population dynamics during reaching. Nature. https://doi.org/10.1038/nature11129 7. Cichy, R. M. et al. (2012) Imagery and perception share cortical representations of content and location. Cereb. Cortex 22, 372 -- 380 8. Cox, D. D. and Savoy, R. L. (2003) Functional magnetic resonance imaging (fMRI) "brain reading": detecting and classifying distributed patterns of fMRI activity in human visual cortex. NeuroImage 19, 261 -- 270 9. Chung, S., Lee, D. D., & Sompolinsky, H. (2016). Linear readout of object manifolds. Physical Review E, 93(6), 060301. 10. Chung, S., Lee, D. D., & Sompolinsky, H. (2018). Classification and geometry of general perceptual manifolds. Physical Review X, 8(3), 031003. 11. Decharms, R. C., & Zador, A. (2000). Neural representation and the cortical code. Annual review of neuroscience, 23(1), 613-647. 12. Dennett, D. (1987) The Intentional Stance, MIT Press 13. de-Wit, L., Alexander, D., Ekroll, V., & Wagemans, J. (2016). Is neuroimaging measuring information in the brain? Psychonomic bulletin & review, 23(5), 1415-1428. 14. DiCarlo, J. J. and Cox, D. (2007) Untangling invariant object recognition. Trends Cogn. Sci. 11, 334-341 15. DiCarlo, J. J., Zoccolan, D., & Rust, N. C. (2012). How does the brain solve visual object recognition?. Neuron, 73(3), 415-434. 16. Diedrichsen, J. (2018). Representational Models. In The Cognitive Neurosciences (M.S. Gazzaniga (Ed.)). 17. Diedrichsen, J., & Kriegeskorte, N. (2017). Representational models: A common framework for understanding encoding, pattern-component, and representational-similarity analysis. PLOS Computational Biology, 13(4), e1005508. https://doi.org/10.1371/journal.pcbi.1005508 • This paper explains the conceptual and mathematical relationships between linear encoding models, pattern-component modeling, and representational similarity analysis and 18 how these methods provide complementary tools in a single toolbox for representational analyses. 18. Diedrichsen, J., Ridgway, G. R., Friston, K. J., & Wiestler, T. (2011). Comparing the similarity and spatial structure of neural representations: A pattern-component model. NeuroImage, 55(4), 1665 -- 1678. https://doi.org/10.1016/j.neuroimage.2011.01.044 • This paper introduces pattern component modeling. 19. Diedrichsen, J., Wiestler, T., & Krakauer, J. W. (2012). Two distinct ipsilateral cortical representations for individuated finger movements. Cerebral Cortex, 23(6), 1362-1377. 20. Duda, R. O., Hart, P. E., & Stork, D. G. (2012). Pattern classification. John Wiley & Sons. 21. Dumoulin, S. O., & Wandell, B. A. (2008). Population receptive field estimates in human visual cortex. NeuroImage, 39(2), 647 -- 660. https://doi.org/10.1016/j.neuroimage.2007.09.034 • This paper introduces nonlinear encoding models for early visual representations, which enabled more rigorous investigations of early vision with neuroimaging data. 22. Eickenberg, M., Gramfort, A., Varoquaux, G., & Thirion, B. (2017). Seeing it all: Convolutional network layers map the function of the human visual system. NeuroImage, 152, 184-194. 23. Felleman, D. J., & Van Essen, D. (1991). Distributed hierarchical processing in the primate cerebral cortex. Cerebral cortex (New York, NY: 1991), 1(1), 1-47. 24. Freedman, D. J., Riesenhuber, M., Poggio, T., & Miller, E. K. (2002). Visual Categorization and the Primate Prefrontal Cortex: Neurophysiology and Behavior. Journal of Neurophysiology, 88(2), 929 -- 941. https://doi.org/10.1152/jn.2002.88.2.929 25. Friston, K. J., Holmes, A. P., Worsley, K. J., Poline, J. P., Frith, C. D., & Frackowiak, R. S. (1994). Statistical parametric maps in functional imaging: a general linear approach. Human brain mapping, 2(4), 189-210. 26. Goodfellow, I., Pouget-Abadie, J., Mirza, M., Xu, B., Warde-Farley, D., Ozair, S., ... & Bengio, Y. (2014). Generative adversarial nets. In Advances in neural information processing systems (pp. 2672-2680). 27. Hastie, T., Tibshirani, R., & Friedman, J. (2009). The Elements of Statistical Learning. New York, NY: Springer New York. https://doi.org/10.1007/978-0-387-84858-7 28. Haufe, S., Meinecke, F., Görgen, K., Dähne, S., Haynes, J.-D., Blankertz, B., & Biessmann, F. (2014). On the interpretation of weight vectors of linear models in multivariate neuroimaging. NeuroImage, 87, 96 -- 110. https://doi.org/10.1016/j.neuroimage.2013.10.067 • This paper provides a careful mathematical discussion of the difficulty of interpreting the weights of linear models. 29. Haxby, J. V. et al. (2001). Distributed and Overlapping Representations of Faces and Objects in Ventral Temporal Cortex. Science, 293(5539), 2425 -- 2430. https://doi.org/10.1126/science.1063736 • A seminal decoding study of ventral-stream category representations that helped the field conceptualize neuroimaging data as reflections of distributed neuronal population codes. 30. Haynes, J.-D. (2015). A Primer on Pattern-Based Approaches to fMRI: Principles, Pitfalls, and Perspectives. Neuron, 87(2), 257 -- 270. https://doi.org/10.1016/j.neuron.2015.05.025 31. Haynes, J.-D., & Rees, G. (2006). Decoding mental states from brain activity in humans. Nature Reviews Neuroscience, 7(7), 523 -- 534. https://doi.org/10.1038/nrn1931 19 32. Hebart, M. N., & Baker, C. I. (2017). Deconstructing multivariate decoding for the study of brain function. Neuroimage. 33. Hebart, M. N., Görgen, K., & Haynes, J. D. (2015). The Decoding Toolbox (TDT): a versatile software package for multivariate analyses of functional imaging data. Frontiers in neuroinformatics, 8, 88. 34. Hong, H., Yamins, D. L. K., Majaj, N. J., & DiCarlo, J. J. (2016). Explicit information for category-orthogonal object properties increases along the ventral stream. Nature Neuroscience, 19(4), 613 -- 622. https://doi.org/10.1038/nn.4247 35. Huth, A. G. et al. (2012) A continuous semantic space describes the representation of thousands of object and action categories across the human brain. Neuron 76, 1210 -- 1224 36. Kamitani, Y., & Tong, F. (2005). Decoding the visual and subjective contents of the human brain. Nature Neuroscience, 8(5), 679 -- 685. https://doi.org/10.1038/nn1444 • A seminal decoding study of primary visual representations of oriented gratings, which established that functional magnetic resonance imaging enables orientation decoding. 37. Kay, K. N., Naselaris, T., Prenger, R. J., & Gallant, J. L. (2008). Identifying natural images from human brain activity. Nature, 452(7185), 352 -- 355. https://doi.org/10.1038/nature06713 • A seminal encoding and decoding study that helped the field envision how neuroimaging data can be used to test image-computable vision models. 38. Kell, A. J. E., Yamins, D. L. K., Shook, E. N., Norman-Haignere, S. V., & McDermott, J. H. (2018). A Task-Optimized Neural Network Replicates Human Auditory Behavior, Predicts Brain Responses, and Reveals a Cortical Processing Hierarchy. Neuron, 98(3), 630- 644.e16. https://doi.org/10.1016/j.neuron.2018.03.044 39. Khaligh-Razavi, S.-M., Henriksson, L., Kay, K., & Kriegeskorte, N. (2016). Fixed versus mixed RSA: Explaining visual representations by fixed and mixed feature sets from shallow and deep computational models. Journal of Mathematical Psychology. https://doi.org/10.1016/j.jmp.2016.10.007 40. King, J. R., & Dehaene, S. (2014). Characterizing the dynamics of mental representations: the temporal generalization method. Trends in cognitive sciences, 18(4), 203-210. 41. Kriegeskorte, N. (2011). Pattern-information analysis: From stimulus decoding to computational-model testing. NeuroImage, 56(2), 411 -- 421. https://doi.org/10.1016/j.neuroimage.2011.01.061 42. Kriegeskorte, N. (2015). Crossvalidation in brain imaging analysis (No. biorxiv;017418v1). Retrieved from http://biorxiv.org/lookup/doi/10.1101/017418 43. Kriegeskorte, N. (2015). Deep Neural Networks: A New Framework for Modeling Biological Vision and Brain Information Processing. Annual Review of Vision Science 1, 417 -- 446. • A review and perspective paper developing a framework for how to build neural network models of brain information processing with constraints from brain-activity data. 44. Kriegeskorte, N., & Bandettini, P. (2007). Analyzing for information, not activation, to exploit high-resolution fMRI. Neuroimage, 38(4), 649-662. 45. Kriegeskorte, N. et al. (2008) Representational similarity analysis -- connecting the branches of systems neuroscience. Front. Syst. Neurosci. 2, 4. • The paper introducing representational similarity analysis. 46. Kriegeskorte, N. et al. (2007) Individual faces elicit distinct response patterns in human anterior temporal cortex. Proc. Natl Acad. Sci. U. S. A. 104, 20600-20605 20 47. Kriegeskorte, N., & Diedrichsen, J. (2016). Inferring brain-computational mechanisms with models of activity measurements. Philosophical Transactions of the Royal Society B: Biological Sciences, 371(1705), 20160278. https://doi.org/10.1098/rstb.2016.0278 48. Kriegeskorte, N., Goebel, R., & Bandettini, P. (2006). Information-based functional brain mapping. Proceedings of the National Academy of Sciences of the United States of America, 103(10), 3863 -- 3868. https://doi.org/10.1073/pnas.0600244103 49. Majaj, N. J., Hong, H., Solomon, E. A., & DiCarlo, J. J. (2015). Simple learned weighted sums of inferior temporal neuronal firing rates accurately predict human core object recognition performance. Journal of Neuroscience, 35(39), 13402-13418. 50. Mesgarani, N., David, S. V., Fritz, J. B., & Shamma, S. A. (2009). Influence of context and behavior on stimulus reconstruction from neural activity in primary auditory cortex. Journal of neurophysiology, 102(6), 3329-3339. 51. Mitchell, T. M. et al. (2008) Predicting human brain activity associated with the meanings of nouns. Science 320, 1191 -- 1195 • A seminal encoding study of semantic brain representations using semantic property descriptions to predict brain representations of novel stimuli. 52. Miyawaki, Y., Uchida, H., Yamashita, O., Sato, M., Morito, Y., Tanabe, H. C., … Kamitani, Y. (2008). Visual Image Reconstruction from Human Brain Activity using a Combination of Multiscale Local Image Decoders. Neuron, 60(5), 915 -- 929. https://doi.org/10.1016/j.neuron.2008.11.004 53. Mur, M., Bandettini, P. A., & Kriegeskorte, N. (2009). Revealing representational content with pattern-information fMRI -- an introductory guide. Social Cognitive and Affective Neuroscience, 4(1), 101 -- 109. https://doi.org/10.1093/scan/nsn044 54. Naselaris, T., & Kay, K. N. (2015). Resolving Ambiguities of MVPA Using Explicit Models of Representation. Trends in Cognitive Sciences, 19(10), 551 -- 554. https://doi.org/10.1016/j.tics.2015.07.005 55. Naselaris, T., Kay, K. N., Nishimoto, S., & Gallant, J. L. (2011). Encoding and decoding in fMRI. NeuroImage, 56(2), 400 -- 410. https://doi.org/10.1016/j.neuroimage.2010.07.073 56. Naselaris, T., Prenger, R. J., Kay, K. N., Oliver, M., & Gallant, J. L. (2009). Bayesian Reconstruction of Natural Images from Human Brain Activity. Neuron, 63(6), 902 -- 915. https://doi.org/10.1016/j.neuron.2009.09.006 57. Nili, H., Wingfield, C., Walther, A., Su, L., Marslen-Wilson, W., & Kriegeskorte, N. (2014). A toolbox for representational similarity analysis. PLoS computational biology, 10(4), e1003553. 58. Norman, K. A. et al. (2006) Beyond mindreading: multi-voxel pattern analysis of fMRI data. Trends Cogn. Sci. 10, 424 -- 430; 59. Nunez-Elizalde, A. O., Huth, A. G., & Gallant, J. L. (2018). Voxelwise encoding models with non-spherical multivariate normal priors. bioRxiv, 386318. 60. Paninski, L., & Cunningham, J. (2017). Neural data science: accelerating the experiment- analysis-theory cycle in large-scale neuroscience. bioRxiv, 196949. 61. Paninski, L., Pillow, J., & Lewi, J. (2007). Statistical models for neural encoding, decoding, and optimal stimulus design. Progress in brain research, 165, 493-507. 21 62. Parthasarathy, N., Batty, E., Falcon, W., Rutten, T., Rajpal, M., Chichilnisky, E. J., & Paninski, L. (2017). Neural networks for efficient bayesian decoding of natural images from retinal neurons. In Advances in Neural Information Processing Systems (pp. 6434-6445). 63. Pereira, F., Mitchell, T., & Botvinick, M. (2009). Machine learning classifiers and fMRI: a tutorial overview. NeuroImage, 45(1 Suppl), S199-209. https://doi.org/10.1016/j.neuroimage.2008.11.007 64. Pillow, J. W., Ahmadian, Y., & Paninski, L. (2011). Model-based decoding, information estimation, and change-point detection techniques for multineuron spike trains. Neural computation, 23(1), 1-45. 65. Poggio, T., & Girosi, F. (1990). Networks for approximation and learning. Proceedings of the IEEE, 78(9), 1481-1497. 66. Raizada, R. D., & Kriegeskorte, N. (2010). Pattern‐information fMRI: New questions which it opens up and challenges which face it. International Journal of Imaging Systems and Technology, 20(1), 31-41. 67. Ritchie, J. B., Kaplan, D. M., & Klein, C. (2017). Decoding the brain: Neural representation and the limits of multivariate pattern analysis in cognitive neuroscience. The British Journal for the Philosophy of Science. 68. Santoro, R., Moerel, M., De Martino, F., Valente, G., Ugurbil, K., Yacoub, E., & Formisano, E. (2017). Reconstructing the spectrotemporal modulations of real-life sounds from fMRI response patterns. Proceedings of the National Academy of Sciences, 114(18), 4799-4804. 69. Shadlen, M. N., Britten, K. H., Newsome, W. T., & Movshon, J. A. (1996). A computational analysis of the relationship between neuronal and behavioral responses to visual motion. Journal of Neuroscience, 16(4), 1486-1510. 70. Simoncelli, E. P., & Olshausen, B. A. (2001). Natural image statistics and neural representation. Annual review of neuroscience, 24(1), 1193-1216. 71. Stokes, M. et al. (2009) Top-down activation of shape-specific population codes in visual cortex during mental imagery. J Neurosci. 29, 1565-1572 72. Taylor, J., & Tibshirani, R. J. (2015). Statistical learning and selective inference. Proceedings of the National Academy of Sciences, 112(25), 7629-7634. 73. Thirion, B., Duchesnay, E., Hubbard, E., Dubois, J., Poline, J. B., Lebihan, D., & Dehaene, S. (2006). Inverse retinotopy: inferring the visual content of images from brain activation patterns. Neuroimage, 33(4), 1104-1116. 74. Tishby, N., Pereira, F. C., & Bialek, W. (2000). The information bottleneck method. arXiv preprint physics/0004057. 75. Tong, F. and Pratte, M. S. (2012) Decoding patterns of human brain activity. Ann. Rev. Psychol. 63, 483-509 76. van Gerven, M. A. (2017). A primer on encoding models in sensory neuroscience. Journal of Mathematical Psychology, 76, 172-183. 77. Varoquaux, G., Raamana, P. R., Engemann, D. A., Hoyos-Idrobo, A., Schwartz, Y., & Thirion, B. (2017). Assessing and tuning brain decoders: cross-validation, caveats, and guidelines. NeuroImage, 145, 166-179. 78. Walther, A., Nili, H., Ejaz, N., Alink, A., Kriegeskorte, N., & Diedrichsen, J. (2016). Reliability of dissimilarity measures for multi-voxel pattern analysis. Neuroimage, 137, 188-200. 22 79. Walther, D. B. et al. (2012) To err is human: correlating fMRI decoding and behavioral errors to probe the neural representation of natural scene categories. In: Visual Population Codes - toward a Common Multivariate Framework for Cell Recording and Functional Imaging, (Kriegeskorte, N., Kreiman, G. eds.), pp. 391-416, MIT Press 80. Weichwald, S., Meyer, T., Özdenizci, O., Schölkopf, B., Ball, T., & Grosse-Wentrup, M. (2015). Causal interpretation rules for encoding and decoding models in neuroimaging. Neuroimage, 110, 48-59. • This paper introduces a set of causal interpretation rules for encoding and decoding models on the basis of the technical literature on causal inference. 81. Wen, H., Shi, J., Chen, W., & Liu, Z. (2018). Deep Residual Network Predicts Cortical Representation and Organization of Visual Features for Rapid Categorization. Scientific reports, 8(1), 3752. 82. Williams, M. A., Dang, S., & Kanwisher, N. G. (2007). Only some spatial patterns of fMRI response are read out in task performance. Nature neuroscience, 10(6), 685. 83. Wu, M. C. K., David, S. V., & Gallant, J. L. (2006). Complete functional characterization of sensory neurons by system identification. Annu. Rev. Neurosci., 29, 477-505. 84. Yamane, Y. et al. (2008) A neural code for three-dimensional object shape in macaque inferotemporal cortex. Nat. Neurosci. 11, 1352-1360 85. Yamins, D. L. K. & DiCarlo, J. J. Using goal-driven deep learning models to understand sensory cortex. Nature Neuroscience 19, 356 -- 365 (2016). • A review and perspective paper describing a framework for modeling sensory processing by training deep neural networks on tasks and testing them with brain-activity data. 86. Yamins, D. L. K. et al. Performance-optimized hierarchical models predict neural responses in higher visual cortex. Proceedings of the National Academy of Sciences 111, 8619 -- 8624 (2014). 87. Young, M. P. (1992). Objective analysis of the topological organization of the primate cortical visual system. Nature, 358(6382), 152. 23
1104.5458
1
1104
2011-04-28T17:56:32
An MRI-Derived Definition of MCI-to-AD Conversion for Long-Term, Automati c Prognosis of MCI Patients
[ "q-bio.NC", "physics.med-ph" ]
Alzheimer's disease (AD) and mild cognitive impairment (MCI), continue to be widely studied. While there is no consensus on whether MCIs actually "convert" to AD, the more important question is not whether MCIs convert, but what is the best such definition. We focus on automatic prognostication, nominally using only a baseline image brain scan, of whether an MCI individual will convert to AD within a multi-year period following the initial clinical visit. This is in fact not a traditional supervised learning problem since, in ADNI, there are no definitive labeled examples of MCI conversion. Prior works have defined MCI subclasses based on whether or not clinical/cognitive scores such as CDR significantly change from baseline. There are concerns with these definitions, however, since e.g. most MCIs (and ADs) do not change from a baseline CDR=0.5, even while physiological changes may be occurring. These works ignore rich phenotypical information in an MCI patient's brain scan and labeled AD and Control examples, in defining conversion. We propose an innovative conversion definition, wherein an MCI patient is declared to be a converter if any of the patient's brain scans (at follow-up visits) are classified "AD" by an (accurately-designed) Control-AD classifier. This novel definition bootstraps the design of a second classifier, specifically trained to predict whether or not MCIs will convert. This second classifier thus predicts whether an AD-Control classifier will predict that a patient has AD. Our results demonstrate this new definition leads not only to much higher prognostic accuracy than by-CDR conversion, but also to subpopulations much more consistent with known AD brain region biomarkers. We also identify key prognostic region biomarkers, essential for accurately discriminating the converter and nonconverter groups.
q-bio.NC
q-bio
An MRI-Derived Definition of MCI-to-AD Conversion for Long-Term, Automatic Prognosis of MCI Patients for the Alzheimer's Disease Neuroimaging Initiative $ Yaman Aksua, David J. Millerb, George Kesidisb,c, Don C. Biglerd, Qing X. Yanga aCenter for NMR Research, Department of Radiology, Penn State University College of bElectrical Engineering Department, Penn State University, University Park, PA, USA cComputer Science and Engineering Department, Penn State University, University Medicine, Hershey, PA, USA dCenter for Emerging Neurotechnology and Imaging, Penn State University College of Medicine, Hershey, PA, USA Park, PA, USA Abstract Alzheimer's disease (AD), and its precursor state, mild cognitive impairment (MCI), continue to be widely studied. While there is no consensus on whether MCIs actually "convert" to AD, this concept is widely applied, allowing sta- tistical testing and machine learning methods to help identify early disease biomarkers and build models for predicting disease progression. Thus, the more important question is not whether MCIs convert, but what is the best such definition. We focus on automatic prognostication, nominally using only a baseline image brain scan, of whether an MCI individual will convert to AD within a multi-year period following the initial clinical visit. This is in fact not a traditional supervised learning problem since, in ADNI, there are no definitive labeled examples of MCI conversion. It is not truly unsupervised, either, since there are (labeled) AD and Control subjects, as well as clini- cal and cognitive scores for MCIs. Prior works have defined MCI subclasses $Data used in preparation of this article were obtained from the Alzheimer's Dis- ease Neuroimaging Initiative (ADNI) database (adni.loni.ucla.edu). As such, the in- vestigators within the ADNI contributed to the design and implementation of ADNI and/or provided data but did not participate in analysis or writing of this report. A complete listing of ADNI investigators can be found at: http://adni.loni.ucla.edu/wp- content/uploads/how to apply/ADNI Authorship List.pdf August 7, 2018 based on whether or not clinical/cognitive scores significantly change from baseline. There are serious concerns with these definitions, however, since e.g. most MCIs (and ADs) do not change from a baseline CDR=0.5 at any subsequent visit in ADNI, even while physiological changes may be occur- ring. These works ignore rich phenotypical information in an MCI patient's brain scan and labeled AD and Control examples, in defining conversion. We propose an innovative conversion definition, wherein an MCI patient is declared to be a converter if any of the patient's brain scans (at follow-up visits) are classified "AD" by an (accurately-designed) Control-AD classifier. This novel definition bootstraps the design of a second classifier, specifically trained to predict whether or not MCIs will convert. This second classifier thus predicts whether an AD-Control classifier will predict that a patient has AD. Our results demonstrate this new definition leads not only to much higher prognostic accuracy than by-CDR conversion, but also to subpopu- lations much more consistent with known AD brain region biomarkers. We also identify key prognostic region biomarkers, essential for accurately dis- criminating the converter and nonconverter groups. Keywords: Alzheimer's, mild cognitive impairment, AD conversion, MRI, support vector machines, feature selection, AD biomarkers, MFE, RFE 1. Introduction The dementing illness Alzheimer's disease (AD), and the transitional state between normal aging and AD referred to as mild cognitive impair- ment (MCI) continue to be widely studied. Individuals diagnosed with MCI have memory impairment, yet without meeting dementia criteria. Annually ≈ 10-15% of people with MCI are diagnosed with AD (Petersen, 2004). More- over, prior to symptom onset, brain abnormalities have been found in people with MCI, as ascertained by retroactive evaluation of longitudinal MRI scans (Davatzikos et al., 2008). There is no consensus on whether MCI patients actually "convert" to AD. First, some MCI patients may suffer from other neurodegenerative disorders (e.g., Lewy body dementia, vascular dementia and/or frontotemporal dementia). Second, it is possible that all other MCI patients already have AD, but at a preclinical stage. AD diagnosis itself may not be considered definitive without e.g. confirming pathologies such as the amyloid deposits detectable at autopsy. Regardless of whether MCI patients truly "convert" to AD or not, the concept of MCI-to-AD conversion has been 2 widely applied, e.g. (Chou et al., 2009; Davatzikos et al., 2010; Misra et al., 2009; Vemuri et al., 2009; Zhang et al., 2011) and is utilitarian -- defining MCI (converter and nonconverter) subgroups allows use of statistical group difference tests and machine learning methods to help identify early disease biomarkers and to build models for predicting disease progression. For these purposes, the more important question is not whether MCIs convert, but rather what is the best such definition. Accordingly, here we focus on the following Aim: automatic prognos- tication, (nominally) using only a baseline brain scan, of whether an MCI individual will convert to AD within a multi-year (three year) period follow- ing an initial (baseline) clinical visit1. While only image voxel-based features are evaluated here for use by our classifier, our framework is extensible to incorporating other baseline clinical information (e.g. weight, gender, education level, genetic information, and clinical cognitive scores such as the Mini Mental State Exam (MMSE)) into the decisionmaking. Moreover, our approach can also incorporate the recent, promising cerebrospinal fluid (CSF) based markers (De Meyer et al., 2010). However, as this requires an invasive spinal tap, we focus here on image scans, which are routinely performed for subjects with MCI. We do not hypothesize that, within ADNI, there are actually two sub- classes of MCI subjects when evaluated over the very long term -- those that (eventually) convert to AD, and those that do not. Even if an overwhelm- ing majority of MCI subjects will eventually convert, identifying the sub- group likely to convert within several years has several compelling utilities: 1) early prognosis, to assist family planning; 2) facilitating group-targeted treatments/drug trials; 3) we identify key prognostic brain "biomarker" re- gions, i.e. those found to be most critical for accurately discriminating our "converter" and "nonconverter" groups. These regions may shed light on disease etiology. Distinguishing AD converters from nonconverters is a binary (two-class) classification problem. Moreover, it may appear this classification problem can be directly addressed via supervised learning methods (Duda et al., 2001). However, it is in fact an unconventional problem, lying somewhere between su- pervised classification and unsupervised classification (clustering), and thus 1Our system performs three-year ahead prediction because it is designed based on the ADNI database, which followed participants for a period of up to three years. 3 requiring a unique approach. To appreciate this, consider the ADNI cohort of MCI individuals. ADNI consists of clinical information and image scans on hundreds of participants, taken at six-month intervals for up to three years. A clinical label (AD, MCI, or Control) was assigned to each participant at first visit2. Even though a probable AD definition based on CDR and MMSE scores and NINCDS/ADRDA criteria has been used, e.g. (Leow et al., 2009; Zhang et al., 2011), to provide follow-up assessment for MCI patients, this is strictly a clinically driven definition, based on a clinical rating (CDR) and a cognitive score (MMSE) whose difficulties will be pointed out shortly. This is not a definitive (autopsy-based) determination of AD, nor is it a definition based on physiological brain changes. Even if the probable AD definition has very high specificity, it may not be sufficiently sensitive, i.e. there may be pa- tients who are undergoing significant physiological brain changes consistent with conversion, yet without clinical manifestation. Accordingly, we will approach the conversion problem from a perspective as agnostic and unbiased as possible, and simply state that it is not defini- tively known which MCI participants in ADNI truly converted to AD within three years. In conventional supervised classifier learning, one has labeled training examples, used for designing the classifier, and labeled test exam- ples, used to estimate the classifier's generalization accuracy. For predicting whether MCI participants in ADNI convert to AD, we in fact have neither. Thus, our problem is not conventional supervised learning. On the other hand, consider unsupervised clustering (Duda et al., 2001). Here, even if one knows the number of clusters (classes) present, there is no prior knowledge on what is a good clustering -- one is simply looking for underlying group- ing tendency in the data. Clearly, our problem does not fit unsupervised clustering, either -- while we have no labeled MCI converter/nonconverter instances per se, 1) there are two designated classes of interest (converter and nonconverter); and 2) there are known class characteristics -- conversion to AD should, plausibly, mean that: (i) a clinical measure such as CDR or a cognitive measure has changed and/or (ii) there are changes in brain fea- tures more characteristic of AD subjects than normal/healthy subjects. Note that in ADNI we do have plentiful labeled AD and normal/healthy (Control) 2Clinicians derive the AD/MCI/Control label based on multiple criteria, which may include Clinical Dementia Rating (CDR), whose possible values are: 0=none, 0.5=ques- tionable, 1=mild, 2=moderate, 3=severe. 4 examples to help assess ii). Based on the above, MCI prognosis is an interesting and novel problem, lying somewhere between supervised and unsupervised classification. The crux of this problem is to craft criteria through which meaningful MCI sub- groups can be defined, well-capturing notions of "AD converter" and "non- converter". To help guide development and evaluation of candidate defini- tions, we state the following three desiderata: 1) The proposed definition of AD converter should be plausible and should exploit the available, rele- vant information in the ADNI database (e.g. image data, labeled AD and Control examples, and clinical information). To appreciate 1), note that the MCI population could be dichotomized in many ways, e.g. by height, and there might be significant clustering tendency with respect to height, but such a grouping is likely meaningless for MCI prognosis; 2) A classifier trained based on these class definitions should generalize well on test data (not used for training the classifier) -- this quantifies how accurately we can discriminate the classes that we have defined. If we create what we believe to be good definitions, but ones that cannot be accurately discriminated, that would not be useful clinically; 3) The class definitions should be vali- dated using known AD conversion biomarkers (i.e., external measures) such as measured changes from baseline both in volumes of brain regions known to be associated with the disease (Schuff et al., 2010) and in cognitive test scores (such as the clinical MMSE measure). Prior Related Work Several prior works, e.g. (Davatzikos et al., 2010; Misra et al., 2009; Wang et al., 2010), defined converter and nonconverter classes solely according to whether the baseline visit CDR score of 0.5 rose or stayed the same over all visits. Change in CDR has also been used as surrogate ground-truth for cognitive decline in a number of other papers, e.g. (Chou et al., 2010; Vemuri et al., 2009). While CDR gives a workable conversion definition, it should be evaluated with respect to the three desiderata above. We will evaluate 2) and 3) in the sequel. With respect to 1), one should challenge a CDR-based conversion definition. First, CDR is not an effective discriminator between the AD and MCI groups, i.e. there is very significant AD-MCI overlap, not only with respect to CDR=1 but even 0.5 -- for ADNI, the majority of the hundreds of AD subjects used in our experiments start (at first visit) at CDR=0.5 and stay at 0.5 at all later visits; likewise, nearly all MCI subjects start at 0.5, with a large majority of these also staying at 0.5 for all visits. This latter fact further implies difficulties in finding an adequate 5 number of conversion-by-CDR subjects in ADNI, both for accurate classifier training and test set evaluation. For the even more stringent probable AD definition (meeting MMSE and NINCDS/ADRDA criteria, in addition to CDR changing from 0.5 to 1) there are necessarily even fewer MCI converters for classifier training and testing. Second, a purely CDR-based (or "probable AD" based) conversion definition ignores the (rich) phenotypical information in an MCI subject's image brain scans and does not exploit the labeled AD and normal/healthy (Control) examples in ADNI. These prior works do treat features derived from brain scans as the covariates (the inputs) to the classifier/predictor. However, we believe the MCI brain scans can themselves be used, in conjunction with the labeled AD and Control examples, to help define more accurate surrogate ground-truth. Previous work has demonstrated that structural MRI analysis is useful for identifying AD biomarkers in individual brain regions (Chetelat et al., 2002; Fan et al., 2008; Fennema-Notestine et al., 2009) -- e.g., cortical thin- ning (Lerch et al., 2005; Thompson et al., 2003), ventricle dilation and gaping (Chou et al., 2010, 2009; Schott et al., 2005), volumetric and shape changes in the hippocampus and entorhinal cortex (Csernansky et al., 2005; de Leon et al., 2006; Stoub et al., 2005), and temporal lobe shrinkage (Rusinek et al., 2004). It is important to capture interaction effects across multiple brain regions. (Davatzikos et al., 2008; Misra et al., 2009; Vemuri et al., 2008; Wang et al., 2010) did jointly analyze voxels (or regions) spanning the entire brain and did build classifiers or predictors. Moreover, as part of their work, (Wang et al., 2010) investigated prediction of future decline in MCI subjects working from baseline MRI scans, which is the primary subject of our cur- rent paper. However, there are several limitations of these past works. First, all these studies used the previously discussed CDR and cognitive measures such as MMSE, which has been described as noisy and unreliable, as the ground-truth prediction targets for classifier/regressor training. In (Chou et al., 2010), the authors state: "Cognitive assessments are notoriously variable over time, and there is increasing evidence that neuroimaging may provide accurate, reproducible measures of brain atrophy." Even in (Wang et al., 2010), where MMSE was treated as the measure of decline and the ground- truth regression target, the authors acknowledged that "individual cognitive evaluations are known to be extremely unstable and depend on a number of factors unrelated to...brain pathology." Such factors include sleep depri- vation, depression, other medical conditions, and medications. Even though MMSE is widely used by clinicians, these comments (even if not universally 6 accepted), do indicate MMSE by itself may not be so reliable in quantifying the disease state. Moreover, while (Wang et al., 2010) did build predictors of future MMSE scores working from baseline scans, this was not a main focus of their paper -- their paper focused on predicting the current score. Their prognostic experiments involved a very small sample size (just 26 participants from the ADNI database). Accordingly, it is difficult to draw definitive con- clusions about the accuracy of their prognostic model and their associated brain biomarkers. The main reason the authors chose such a small sample was, as the authors state: "A large part of...ADNI...are from patients who did not display significant cognitive decline...[these] would overwhelm the re- gression algorithm if..used in the...experiment." While this statement (with cognitive decline measured according to MMSE) may be true, that does not mean many of those excluded ADNI subjects are not experiencing significant physiological brain changes/atrophy. The novel approach we next sketch is well-suited to identifying MCI subjects undergoing such changes. Our Neuroimaging-Driven, Trajectory-Based Approach Here, we propose a novel approach for prognosticating putative conver- sion to AD driven by image-based information (and exploiting AD-Control examples), rather than by a single, non-image-based, weakly discriminat- ing clinical measurement such as CDR. Our solution strategy is as follows. We first build an accurate image-based Control-AD classifier (i.e., using AD and Control subjects, we build a Support Vector Machine (SVM) classifier) (Vapnik, 1998). We then apply this classifier to a training population of MCI subjects -- separately, for each subject visit, we determine whether the subject's image is on the AD side or the Control side of the SVM's fixed (hyperplane) decision boundary. In addition to a binary decision, the SVM gives a "score" -- essentially the distance to the classifier's decision boundary. Thus, for each MCI subject, as a function of visits, we get an image-based "phenotypical" score trajectory. We fit a line to each subject's trajectory and extend the line to the sixth visit if the sixth visit is missing. We can then give the following trajectory-based conversion definition: if the extended line either starts on the AD side or crosses to the AD side over the six visits, we declare this person a "converter-by-trajectory". Otherwise, this person is a "nonconverter-by-trajectory"3. In this fashion, we derive ground-truth 3A very small percentage of the MCI population, in our experiments often 1% and not exceeding 5%, may unexpectedly start on the AD side and cross to the Control side. We 7 "converter" and "nonconverter" labels for an (initially unlabeled) training MCI population. These (now) labeled training samples bootstrap the de- sign of a second SVM classifier which uses only the first-visit training set MCI images and is trained to predict whether or not an MCI patient is a "converter-by-trajectory". Essentially, this second (prognostic) classifier is predicting whether, within three years, an AD-Control classifier will predict that a patient has AD. Via these two classifier design steps, we thus build a classification system for our (unconventional) pattern recognition task. SVMs are widely used classifiers whose accuracy is attributed to their maximization of the "margin", i.e. the smallest distance from any training point to the classification boundary. Since the SVM finds a linear discrimi- nant function that maximizes margin, a significant change in score is generally needed to cross from the control side to the AD side, which is thus suggestive of conversion from MCI to AD. This is the premise underlying our approach. The main contributions of our work are: 1) a novel image-based prognos- ticator of MCI-to-AD conversion that we will demonstrate to achieve both better generalization accuracy and much higher correlation with known brain region biomarkers than the CDR-based approach; 2) The finding that MCI subgroups that are strongly correlated with known AD brain region biomark- ers are not so strongly correlated with "cognitive decline" as measured by MMSE; 3) Identification of the brain regions most critical for accurately discriminating between our "converter" and "nonconverter" groups, via ap- plication of margin-based feature selection (MFE) (Aksu et al., 2010) to brain image classification, and demonstration of MFE's better performance than the well-known RFE method (Guyon et al., 2002) on this domain. treat these individuals as outliers and omit them from our experiments. 8 2. Methods 2.1. Subjects and MRI data We used T1-weighted ADNI images4 that have undergone image correc- tion described at the ADNI website.5 ADNI aims to recruit and follow 800 research participants in the 55-90 age range: approximately 200 elderly Con- trols, 400 people with MCI, and 200 people with AD. The number of Control, MCI, and AD participants in our analysis were ≈180, 300, and 120, respec- tively -- experiment-specific detailed descriptions will be provided in Sec. 3. We processed the T1-weighted images as described in Appendix A, produc- ing new images from which we then obtained the features (next discussed) used by our statistical classifiers. 2.2. Features for classification We chose as features the voxel intensities of a processed6 RAVENS im- age, a type of "volumetric density" image (Davatzikos, 1998; Davatzikos et al., 2001; Goldszal et al., 1998; Shen et al., 2003) that has been validated for voxel-based analysis (Davatzikos et al., 2001) and applied both to AD e.g. (Davatzikos et al., 2010; Misra et al., 2009; Wang et al., 2010) and other studies e.g. (Fan et al., 2007)7. For each of the three processed RAVENS tis- sue maps (gray matter (GM), white matter (WM), and ventricle), to reduce 4 Data used in the preparation of this article were obtained from the Alzheimer's Disease Neuroimaging Initiative (ADNI) database (adni.loni.ucla.edu). The ADNI was launched in 2003 by the National Institute on Aging (NIA), the National Institute of Biomedical Imaging and Bioengineering (NIBIB), the Food and Drug Administration (FDA), private pharmaceutical companies and non-profit organizations, as a $60 million, 5-year public- private partnership. The primary goal of ADNI has been to test whether serial mag- netic resonance imaging (MRI), positron emission tomography (PET), other biological markers, and clinical and neuropsychological assessment can be combined to measure the progression of mild cognitive impairment (MCI) and early Alzheimer's disease (AD). De- termination of sensitive and specific markers of very early AD progression is intended to aid researchers and clinicians to develop new treatments and monitor their effectiveness, as well as lessen the time and cost of clinical trials. 5ADNI image correction steps include Gradwarp, N3, and scaling for gradient drift -- see www.loni.ucla.edu/ADNI/Data/ADNI Data.shtml. 6We describe our processing of RAVENS images in Appendix A. 7Of particular interest, (Davatzikos et al., 2001) supported that voxel-based SPM statis- tical analysis, which we perform herein for comparison with our methods, can be performed on RAVENS images. 9 complexity for subsequent processing, we obtained a subsample by succes- sively skipping five voxels along each of the three dimensions, and took as feature set the union of the three subsampled maps. We will also report results for the case of skipping only two voxels, rather than five. Since high-dimensional nonlinear registration (warping) of all individuals to a common atlas (via HAMMER (Shen et al., 2002)) is applied in pro- ducing our features, they capture both volumetric and morphometric brain characteristics, which is important since individuals with AD/MCI typically exhibit brain atrophy (affecting both volume and shape). 2.3. Classification and feature selection for high-dimensional images A challenge in building classifiers for medical images is the relative paucity of available training samples, compared to the huge dimensionality of the voxel space and, thus, to the number of parameters in the classifier model -- in general, the number of parameters may grow at least linearly with dimen- sionality. In the case of 3D images, this could imply even millions of param- eter values (e.g. one per voxel) need to be determined, based on a training set of only a few hundred patient examples. In such cases, classifier over- fitting is likely, which can degrade generalization (test set) accuracy. Here we will apply a linear discriminant function (LDF) classifier with a built-in mechanism to avoid overfitting and with design complexity that scales well with increasing dimensionality - the support vector machine (SVM) (Vap- nik, 1998). The choice of LDF achieving perfect separation (no classification errors) for a given two-class training set is not unique. The SVM, however, is the unique separating LDF that maximizes the margin, i.e. the minimum distance to the classifier decision boundary, over all training samples. In this sense, the SVM maximizes separation of the two classes. For an SVM, unlike a standard LDF, the number of model parameters is bounded by the number of training samples, rather than being controlled by the feature dimension- ality. Since the number of samples is the much smaller number for medical image domains, in this way the SVM greatly mitigates overfitting. SVMs have achieved excellent classification accuracy for numerous scientific and engineering domains, including medical image analysis, (Aksu et al., 2010; Davatzikos et al., 2010; Guyon et al., 2002). Even though SVMs are effective at mitigating overfitting, generalization accuracy may still be improved in some cases by removing features that contribute little discrimination power. Moreover, even if generalization ac- curacy monotonically improves with increasing feature dimensionality, high 10 complexity (both computation and memory storage) of both classifier design and class decisionmaking may outweigh small gains in accuracy achieved by using a huge number of features. Most importantly here, it is often useful to identify the critical subset of features necessary for achieving accurate classification -- these "markers" may shed light on the underlying disease mechanism. In our case, this will help to identify prognostic brain regions, associated with MCI conversion. Unfortunately, there is a huge number of possible feature subsets, with exhaustive subset evaluation practically prohibited even for modest number of features, M , let alone M ∼ 106. Practical feature selection techniques are thus heuristic, with a large range of tradeoffs between accuracy and complex- ity (Guyon et al., 2003). "Front-end" (or "filtering") methods select features prior to classifier training, based on evaluation of discrimination power for individual features or small feature groups. "Wrapper" methods are gen- erally more reliable, interspersing sequential feature selection and classifier design steps, with features sequentially selected to maximize the current sub- set's joint discrimination power. There are also embedded feature selection methods, e.g. for SVMs, use of (cid:96)1-regularization within the SVM design opti- mization (Fung et al., 2004), in order to find "sparse" weight vector solutions, which effectively eliminate many features. For wrappers, there is greedy for- ward selection, with "informative" features added, backward elimination, which starts from the full set and removes features, and more complex bidi- rectional searches. In our work, due to the high feature dimension, we focus on two backward elimination wrappers that afford practical complexity: i) the widely used recursive feature elimination algorithm (RFE) (Guyon et al., 2002), where at each step one removes the feature with least weight magni- tude in the SVM solution. RFE has been applied before to AD (Davatzikos et al., 2010; Misra et al., 2009; Wang et al., 2010); ii) the recent margin- based feature elimination (MFE) algorithm (Aksu et al., 2010), which uses the same objective function (margin) for feature elimination, one consistent with good generalization, that the SVM uses for classifier training (Vapnik, 1998). MFE was shown in (Aksu et al., 2010) to outperform RFE (Guyon et al., 2002) and to achieve results comparable to embedded feature selection for domains with up to 8,000 features (gene microarray classification). Here we will also find that MFE gives better results than RFE. 11 2.4. An MRI-Derived Alternative to CDR-based MCI-to-AD Conversion In the Introduction, we outlined our two classifier design steps for build- ing an automatic prognosticator for an individual with MCI. In this sec- tion, we elaborate on these two steps and give an illustrative example. Our AD-Control classifier, used in the first step, is discussed in Sec. 2.4.1, and our second classifier, used to discriminate converter-by-trajectory (CT) and nonconverter-by-trajectory (NT) classes, is discussed in Sec. 2.4.2. 2.4.1. AD-Control classifier For the AD training population, we chose individual AD visit images with a CDR score of at least 1. For the Control training population, on the other hand, we only chose initial visits, and only those for participants who stayed at CDR=0 throughout all their visits. Thus, we excluded Controls with "questionable dementia" (i.e., CDR=0.5) at any visit. By these choices, we sought to exclude outlier examples or even possibly any mislabeled examples, recalling that CDR for the majority of both AD and MCI participants is 0.5 throughout all visits. 2.4.2. CT-NT classifier Fig. 1 gives an illustrative example of the phenotypical score trajecto- ries for MCI subject described in the Introduction. A positive score is on the Control side and a negative score is on the AD side -- the x-axis represents the AD-Control SVM's decision boundary. Score vs. age is plotted, with each line segment a trajectory obtained by linearly fitting an individual's phenotype scores (and linearly extrapolating if there are missing visits). Nonconverters- by-CDR (N-CDR) and converters-by-CDR (C-CDR) are illustrated in (a) and (b), respectively. Green and black subjects are those whose fitted tra- jectory stayed on the Control side and AD side, respectively, whereas gray lines are subjects who crossed to the AD side. Thus, by our conversion-by- trajectory definition, the green group is the nonconverters-by-trajectory, and the black and gray groups together are the converters-by-trajectory. Subject counts for these groups are given in the figure legends. The outlier sub- jects are shown in orange -- there are five, making up less than 2% of the MCI cohort. Notice, intriguingly, from the left figure that more than one third of all (non-orange) MCI patients (106 of 298) are converters by tra- jectory and yet nonconverters according to CDR -- i.e., there is a very large percentage of patients for which the two converter definitions disagree, with the neuroimage-based definition indicating disease state changes that are not 12 (a) 13 (b) Figure 1: AD-Control SVM score trajectories for MCI subjects. (a) Nonconverters-by- CDR. (b) Converters-by-CDR. 405060708090100(cid:239)2(cid:239)1.5(cid:239)1(cid:239)0.500.511.52agescore 144 (56.9%)42 (16.6%)64 (25.3%)3 (1.2%)405060708090100(cid:239)3(cid:239)2.5(cid:239)2(cid:239)1.5(cid:239)1(cid:239)0.500.511.5agescore 11 (22.0%)14 (28.0%)23 (46.0%)2 (4.0%) predicted using the clinical, CDR-based definition. Likewise, an additional 3% of all MCI subjects (11 of 298) "defy" their by-CDR converter label in that they do not reach the AD side of the decision boundary. Based on these trajectories, i.e. whether or not the AD side is visited, we derive the ground-truth 'CT' and 'NT' labels for all MCI subjects. We then build a CT-NT classifier using as input only the image scans at initial visit.8 3. Results and Discussion 3.1. Introductory overview In this section we will perform 1) classification experiments to evaluate conversion-by-trajectory and conversion-by-CDR with respect to desidera- tum 2; 2) additional experiments to compare the two definitions with re- spect to desideratum 3; and lastly, 3) experiments to identify prognostic brain "biomarker" regions. It is important to mitigate the potential confounding effect of the subject's age. In our classification experiments, we mitigated in two ways: 1) For every classifier training, each training sample in one class was uniquely paired via "age-matching" with a training sample in the other class (with age separation at most one year). 2) For every linear-kernel SVM classifier, we separately adjusted each feature for age prior to classification using linear fitting. We subtracted the extrapolated line (computed only using "control" samples9) from the feature's value, for all (training and test) samples. As an aside, we note that, given the subsequent linear SVM operation, the representation power of this linear fitting step is essentially equivalent to simply treating age as an additional feature input to the linear SVM classifier. Finally, prior to building classifiers, we normalized feature values to the [0,1] range, which is suitable for the LIBSVM software (Chang et al., 2001) we used for training SVMs. 8For a small percentage of the MCI subjects, we did not obtain the patient's first visit. However, we did ensure that the visit we took as the "initial visit" had a CDR of 0.5. 9For the AD-Control classifier, these are the samples in the Control class. For the CT-NT classifier we computed the line using only the NT samples. 14 (a) By-CDR. (b) By-trajectory. (c) Overlap. Figure 2: A population of 298 MCI subjects in ADNI is shown here, broken up according to the two criteria discussed in Sec. 3.2.1: (a) by-CDR criterion, (b) by-trajectory criterion; with overlap shown in (c). 3.2. Experiments with voxel-based features The test set (generalization) accuracy of the voxel-based AD-Control clas- sifier, built using 70 training samples per class, was 0.89 (86 of 88 Controls, and 16 of 27 AD subjects, were correctly classified.) This classifier, with high specificity for Controls, was then applied to a population of MCI subjects to determine the CT and NT subgroups. 3.2.1. Classification experiments for the MCI population Fig. 2(a) shows the sizes of the converter-by-CDR (C-CDR) and nonconverter- by-CDR (N-CDR) groups within the ADNI MCI cohort for a typical exper- iment in our work. Fig. 2(b) shows the same population broken up as converters-by-trajectory (CT) and nonconverters-by-trajectory (NT). Super- imposing the two charts, Fig. 2(c) illustrates their overlap, where converters by both definitions are accordingly indicated by orange. Since converters- by-CDR are relatively scarce, we used a large majority of them (80%, i.e. 39 individuals among the 48) for the by-CDR classifier's training set, with the rest (20%) put into the test set. We reiterate that a general disadvantage of the by-CDR approach is its scarcity of converter examples -- by contrast, a more balanced number of examples is available for by-trajectory training (at least 100, rather than 39, training samples per class, as in Fig. 2(b)). Note also that if we were to use a "probable AD", rather than a by-CDR converter definition, there would be even fewer converter examples. A fair performance comparison between by-trajectory and by-CDR clas- sification requires: 1) using the same per-class training set size (i.e. 39) for 15 (a)By-CDR.(b)By-trajectory.(c)Overlap.Figure2:Apopulationof298MCIsubjectsinADNI(approximately75%oftheMCIpopulationsizeof400targetedbyADNI)isshownhere,brokenupaccordingtothetwocriteriadiscussedinSec.3.2.1:(a)by-CDRcriterion,(b)by-trajectorycriterion;withoverlapshownin(c).samesets);11ii)wecanmakethetrainingsetsofthetwoclas-sifiersidenticalratherthanmerelysame-sized,aswellasmakethetestsetsidentical.Thislatterapproach,though,willhavesomebiasbecause,inselectingsamplesfortheby-trajectoryclassifier,wewillhavetomakeuseofknowledgeofthesam-ples'conversion-by-CDRstatus(andviceversafortheby-CDRclassifier).Thefirstapproach,ontheotherhand,clearlydoesnothavethisbias.Asbothapproachesarevalidwaysofdealingwithby-CDRdatalimitations,wewillcomparegeneralizationaccuraciesofby-CDRandby-trajectoryclassifiersunderboththesedataselectionschemes,referringtotheseapproachesas"random"and"identical"inthesequel.Ourtraining/testsetselectionprocedureforthe"identicalap-proach"isasfollows.FortheC-CDR-CTgroup(Fig.2(c)),randomlyselect80%ofthegroup(theyellowstripedgroupofsize30inFig.3(a))suchthatacorrespondinggroupwithinN-CDR(whiteportioninFig.2(a))canbefoundthatisbothNTandsatisfiesage-matching.Thiscorrespondinggroupisillus-tratedinFig.3(a)asthewhitestripedgroup(ofsize30),placedoppositefromtheyellowstripedareaitispaired(matched)with.LikewisefortheC-CDR-NTgroup(Fig.2(c)),randomlyselect80%ofthegroup(theyellowstripedgroupofsize9inFig.3(a))suchthatacorrespondinggroupwithinN-CDRcanbefoundthatisbothCTandsatisfiesage-matching.Thiscor-respondinggroupisillustratedinFig.3(a)asthewhitestripedgroup(ofsize9).NoticebycomparingthisfiguretoFig.2(c)thatthetwowhitestripedareasareseparatedbytheCT-NTborder.Wetakethetrainingset -- sharedbytheby-CDRandby-trajectoryclassifiers -- tobepreciselytheunionofthesefourstripedareas.12Subsequently,wetakethetestset -- sharedbythetwoclassifiers -- tobethesubjectswhoareneitherin1)thetrainingset(stripedareas)norin2)thespecialsetofsubjectsshowninsolidgrayinFig.3(a)(alsoshownidenticallyinFig.11Notethatthismeansthatthetrainingsetsfortheconverterandnoncon-verterclassesoftheconversion-by-CDRclassifierarerandomlyselectedfromtheyellowandwhiteregionsinFig.2(a),respectively,withnoconsiderationoftrajectory-based(i.e.red/white)labelingillustratedinFig.2(b).12Fortheby-CDRclassifier,theclassmembershipofanyofthesefoursub-setsofthetrainingsetisillustratedbythecolorbeingyelloworwhiteinFig.3(a).Likewise,fortheby-trajectoryclassifier,classmembershipisillustratedbyredorwhitecolorinFig.3(b).3(b)).13Thatis,thetestsetisthetiledareasinFig.3(a)(or,identically,inFig.3(b)).Noteabovethatsomerandomselectionisbeingemployedinchoosingthetraining/testsetseveninthe"identicalapproach"(whereas,inthe"randomapproach"theselectioniscompletelyrandom).Thus,forbothapproaches,theaccuracyofperfor-mancecomparisonwillbenefitfromaveragingaccuracyresultsovermultipletraining/testsplit"trials".Resultsaveragedacross10trialsaregiveninFig.3(c)foralinear-kernelSVM14;µ±σnotationisusedtoindicatethemeanµandstandarddeviationσofquantitiesacrossthetrials.15Notethatby-trajectory'sgen-eralizationperformanceisashighas0.83,whereasby-CDR'sgeneralizationperformanceisverypoor -- aspoorasrandomguessing(see0.5and0.56tablevalues) -- duemainlytopoorperformanceonnonconverters-by-CDR.Fig.3(d)andFig.3(e)showby-trajectoryandby-CDRresults,respectively,foroneofthe10trials(forthe"randomapproach"),witheachbarindi-catingdistancetotheclassificationboundary16foranMCIsub-jectinatestpopulationofsize88andnonconverters/convertersshowninleft/rightfigures,respectively.Amongthe88subjects,by-trajectorycorrectlyclassified79whereasby-CDRcorrectlyclassifiedonly40.Inaseparateexperiment,weevaluatedusingoneof27sub-samples(ratherthanoneof216subsamples),i.e.essentiallya10-foldincreaseinthenumberof(voxel-based)features,andfoundthattheby-trajectorygeneralizationaccuracyroseto0.91inthe"random"case.Wethentriedbuilding27separateby-converterclassifiers,oneforevery1/27thsubsample(thusef-fectivelyusingthewhole3Dimage),withmajority-basedvot-ingusedtocombinethe27decisions.Thisensembleschemeagainachieved0.91accuracy,i.e.therewasnofurtheraccuracy13Weexcludethis"specialset"(ingray)fromthetestsetsothatallourexperimentsunderthe"identicalapproach"canhaveashared,fixedtestset(forfaircomparisonwitheachother),including,crucially,anexperimentthatwillincludethis"specialset"ofsamplesinthetrainingset.14Forgeneratingallclassificationresultsherein,includingthoseinFig.3(c),weusedSVMclassifiersthatwerebuiltbyemployingthecommonapproachofbootstrap-basedvalidationforselectingtheclassifier's(trial's)hyperparametervalues(Aksuetal.,2010).15µandσofintegerquantities(e.g.samplecounts)areshownroundedup.16Positive/negativedistancemeansnonconverter/convertersideofthebound-ary,respectively.7 both by-CDR and by-trajectory training, and 2) making the test set sizes the same for both classifiers. There are several different ways in which the data can be partitioned into training and test sets, consistent with these two conditions: i) we can perform simple random selection on a class-by-class ba- sis, ensuring only that the two classifiers are given the same training/test set sizes (but not the same sets);10 ii) we can make the training sets of the two classifiers identical rather than merely same-sized, as well as make the test sets identical. This latter approach, though, will have some bias because, in selecting samples for the by-trajectory classifier, we will have to make use of knowledge of the samples' conversion-by-CDR status (and vice versa for the by-CDR classifier). The first approach, on the other hand, clearly does not have this bias. As both approaches are valid ways of dealing with by-CDR data limitations, we will compare generalization accuracies of by-CDR and by-trajectory classifiers under both these data selection schemes, referring to these approaches as "random" and "identical" in the sequel. Our training/test set selection procedure for the "identical approach" is as follows. For the C-CDR-CT group (Fig. 2(c)), randomly select 80% of the group (the yellow striped group of size 30 in Fig. 3(a)) such that a corresponding group within N-CDR (white portion in Fig. 2(a)) can be found that is both NT and satisfies age-matching. This corresponding group is illustrated in Fig. 3(a) as the white striped group (of size 30), placed opposite from the yellow striped area it is paired (matched) with. Likewise for the C-CDR-NT group (Fig. 2(c)), randomly select 80% of the group (the yellow striped group of size 9 in Fig. 3(a)) such that a corresponding group within N-CDR can be found that is both CT and satisfies age-matching. This corresponding group is illustrated in Fig. 3(a) as the white striped group (of size 9). Notice by comparing this figure to Fig. 2(c) that the two white striped areas are separated by the CT-NT border. We take the training set -- shared by the by-CDR and by-trajectory classifiers -- to be precisely the union of these four striped areas.11 Subsequently, we take the test set -- 10Note that this means that the training sets for the converter and nonconverter classes of the conversion-by-CDR classifier are randomly selected from the yellow and white re- gions in Fig. 2(a), respectively, with no consideration of trajectory-based (i.e. red/white) labeling illustrated in Fig. 2(b). 11For the by-CDR classifier, the class membership of any of these four subsets of the training set is illustrated by the color being yellow or white in Fig. 3(a). Likewise, for the by-trajectory classifier, class membership is illustrated by red or white color in Fig. 3(b). 16 shared by the two classifiers -- to be the subjects who are neither in 1) the training set (striped areas) nor in 2) the special set of subjects shown in solid gray in Fig. 3(a) (also shown identically in Fig. 3(b)).12 That is, the test set is the tiled areas in Fig. 3(a) (or, identically, in Fig. 3(b)). Note above that some random selection is being employed in choosing the training/test sets even in the "identical approach" (whereas, in the "random approach" the selection is completely random). Thus, for both approaches, the accuracy of performance comparison will benefit from averaging accuracy results over multiple training/test split "trials". Results averaged across 10 trials are given in Fig. 3(c) for a linear-kernel SVM13; µ± σ notation is used to indicate the mean µ and standard deviation σ of quantities across the trials.14 Note that by-trajectory's generalization performance is as high as 0.83, whereas by-CDR's generalization performance is very poor -- as poor as random guessing (see 0.5 and 0.56 table values) -- due mainly to poor performance on nonconverters-by-CDR. Fig. 3(d) and Fig. 3(e) show by- trajectory and by-CDR results, respectively, for one of the 10 trials (for the "random approach"), with each bar indicating distance to the classification boundary15 for an MCI subject in a test population of size 88 and non- converters/converters shown in left/right figures, respectively. Among the 88 subjects, by-trajectory correctly classified 79 whereas by-CDR correctly classified only 40. Recently, similarly poor by-CDR classification performance was also re- ported in (Davatzikos et al., 2010), where it was found that the majority of (by-CDR) nonconverters "had sharply positive SPARE-AD scores indi- cating significant atrophy similar to AD patients". Since the SPARE-AD score is produced by a classifier that was trained to discriminate Control and AD patients (Fan et al., 2007, 2008), this comment and associated results are consistent both with our conjecture in the Introduction and our above 12We exclude this "special set" (in gray) from the test set so that all our experiments under the "identical approach" can have a shared, fixed test set (for fair comparison with each other), including, crucially, an experiment that will include this "special set" of samples in the training set. 13For generating all classification results herein, including those in Fig. 3(c), we used SVM classifiers that were built by employing the common approach of bootstrap-based validation for selecting the classifier's (trial's) hyperparameter values (Aksu et al., 2010). 14µ and σ of integer quantities (e.g. sample counts) are shown rounded up. 15Positive/negative distance means nonconverter/converter side of the boundary, re- spectively. 17 histogram results, which suggest that there may be a significant number of patients undergoing physiological brain changes consistent with conversion, yet without clinical manifestation. The results above indicate that the conversion-by-CDR definition's two classes are not well-discriminated, and thus, clinical usefulness of this def- inition for our prognostic Aim is expected to be poor. The much greater generalization accuracy of the by-trajectory definition (coupled with its in- herent plausibility as a conversion definition) indicates its greater utility. Increasing the By-Trajectory Image-Based Feature Resolution: In a separate experiment, we evaluated using one of 27 subsamples (rather than one of 216 subsamples), i.e. essentially a 10-fold increase in the number of (voxel-based) features, and found that the by-trajectory generalization ac- curacy rose to 0.91 in the "random" case. We then tried building 27 separate by-converter classifiers, one for every 1/27th subsample (thus effectively us- ing the whole 3D image), with majority-based voting used to combine the 27 decisions. This ensemble scheme again achieved 0.91 accuracy, i.e. there was no further accuracy benefit beyond that from an ≈ 10-fold increase in the number of voxel features. Increasing the By-Trajectory Training Set Size Note that the converter-by-CDR sample scarcity and class-balancing (via age-matching) in the experiments above had the effect of artificially limiting the by-trajectory classifier training set size. Next we investigated how much the generalization accuracy of by-trajectory classification improves when this limitation is removed. The tiled areas in Figures 4(a) and 3(b) are identical, illustrating that in this new experiment (Fig. 4(a)) we used the same test set as previously, for fairness of comparison. However, as indicated by differences in the total striped area between these two charts, we now make the training set much larger than previously. Specifically, for the "identical" case, we used the previous 10 trials but simply augmented a trial's training set with the two large, previously-excluded gray sets16, as these two sets do age-match each other. The results, averaged across the 10 trials, are given in Fig. 4(b). Notice in this figure the now larger per-class training size (on average ≈ 100 rather than 39), and that the random approach uses this size as well. The by-trajectory results in Fig. 4(b) indicate that accuracy improved from 0.78 ± 0.02 (Fig. 3(c)) to 0.84 ± 0.02 for the "identical approach", 16Shown in Fig. 3(b), with size 60. Note that this size can vary from trial to trial. 18 (a) Training/test set selec- tion for by-CDR classifica- tion. (b) Training/test set selec- tion for by-trajectory clas- sification. Sample selection Classifier Converters Test set Count Accuracy Count Accuracy 0.84 ± 0.05 46 ± 1 0.47 ± 0.04 9 ± 1 0.78 ± 0.04 45 ± 1 9 ± 1 0.54 ± 0.04 Nonconverters 45 ± 1 82 ± 2 47 ± 2 82 ± 2 0.81 ± 0.06 0.79 ± 0.07 0.78 ± 0.04 0.74 ± 0.11 Overall accuracy 0.83 ± 0.05 0.50 ± 0.04 0.78 ± 0.02 0.56 ± 0.04 Random By-trajectory approach By-CDR By-trajectory Identical approach By-CDR (c) Average test set classification accuracy using all 11, 293 features, 39±1 per-class training samples. (d) By-trajectory. Left: nonconverters; Right: converters (e) By-CDR. Left: nonconverters; Right: converters Figure 3: Test set accuracy comparison of by-CDR and by-trajectory classification. 19 (a)Training/testsetselectionforby-CDRclassifi-cation.(b)Training/testsetselectionforby-trajectoryclas-sification.SampleClassifierTestsetselectionConvertersNonconvertersOverallCountAccuracyCountAccuracyaccuracyRandomBy-trajectory46±10.81±0.0645±10.84±0.050.83±0.05approachBy-CDR9±10.79±0.0782±20.47±0.040.50±0.04IdenticalBy-trajectory45±10.78±0.0447±20.78±0.040.78±0.02approachBy-CDR9±10.74±0.1182±20.54±0.040.56±0.04(c)Averagetestsetclassificationaccuracyusingall11,293features,39±1per-classtrainingsamples.02040−2.5−2−1.5−1−0.500.511.5 44 test samples [positives: 41, nonpositives: 3]sample indexdistance02040−2.5−2−1.5−1−0.500.511.5 44 test samples [positives: 6, nonpositives: 38]sample indexdistance(d)By-trajectory.Left:nonconverters;Right:con-verters050−1.5−1−0.500.51 79 test samples [positives: 33, nonpositives: 46]sample indexdistance123456789−1.5−1−0.500.51 9 test samples [positives: 2, nonpositives: 7]sample indexdistance(e)By-CDR.Left:nonconverters;Right:convertersFigure3:Testsetaccuracycomparisonofby-CDRandby-trajectoryclassification.9(a)Training/testsetselectionforby-CDRclassifi-cation.(b)Training/testsetselectionforby-trajectoryclas-sification.SampleClassifierTestsetselectionConvertersNonconvertersOverallCountAccuracyCountAccuracyaccuracyRandomBy-trajectory46±10.81±0.0645±10.84±0.050.83±0.05approachBy-CDR9±10.79±0.0782±20.47±0.040.50±0.04IdenticalBy-trajectory45±10.78±0.0447±20.78±0.040.78±0.02approachBy-CDR9±10.74±0.1182±20.54±0.040.56±0.04(c)Averagetestsetclassificationaccuracyusingall11,293features,39±1per-classtrainingsamples.02040−2.5−2−1.5−1−0.500.511.5 44 test samples [positives: 41, nonpositives: 3]sample indexdistance02040−2.5−2−1.5−1−0.500.511.5 44 test samples [positives: 6, nonpositives: 38]sample indexdistance(d)By-trajectory.Left:nonconverters;Right:con-verters050−1.5−1−0.500.51 79 test samples [positives: 33, nonpositives: 46]sample indexdistance123456789−1.5−1−0.500.51 9 test samples [positives: 2, nonpositives: 7]sample indexdistance(e)By-CDR.Left:nonconverters;Right:convertersFigure3:Testsetaccuracycomparisonofby-CDRandby-trajectoryclassification.9(a)Training/testsetselectionforby-CDRclassifi-cation.(b)Training/testsetselectionforby-trajectoryclas-sification.SampleClassifierTestsetselectionConvertersNonconvertersOverallCountAccuracyCountAccuracyaccuracyRandomBy-trajectory46±10.81±0.0645±10.84±0.050.83±0.05approachBy-CDR9±10.79±0.0782±20.47±0.040.50±0.04IdenticalBy-trajectory45±10.78±0.0447±20.78±0.040.78±0.02approachBy-CDR9±10.74±0.1182±20.54±0.040.56±0.04(c)Averagetestsetclassificationaccuracyusingall11,293features,39±1per-classtrainingsamples.02040−2.5−2−1.5−1−0.500.511.5 44 test samples [positives: 41, nonpositives: 3]sample indexdistance02040−2.5−2−1.5−1−0.500.511.5 44 test samples [positives: 6, nonpositives: 38]sample indexdistance(d)By-trajectory.Left:nonconverters;Right:con-verters050−1.5−1−0.500.51 79 test samples [positives: 33, nonpositives: 46]sample indexdistance123456789−1.5−1−0.500.51 9 test samples [positives: 2, nonpositives: 7]sample indexdistance(e)By-CDR.Left:nonconverters;Right:convertersFigure3:Testsetaccuracycomparisonofby-CDRandby-trajectoryclassification.9(a)Training/testsetselectionforby-CDRclassifi-cation.(b)Training/testsetselectionforby-trajectoryclas-sification.SampleClassifierTestsetselectionConvertersNonconvertersOverallCountAccuracyCountAccuracyaccuracyRandomBy-trajectory46±10.81±0.0645±10.84±0.050.83±0.05approachBy-CDR9±10.79±0.0782±20.47±0.040.50±0.04IdenticalBy-trajectory45±10.78±0.0447±20.78±0.040.78±0.02approachBy-CDR9±10.74±0.1182±20.54±0.040.56±0.04(c)Averagetestsetclassificationaccuracyusingall11,293features,39±1per-classtrainingsamples.02040−2.5−2−1.5−1−0.500.511.5 44 test samples [positives: 41, nonpositives: 3]sample indexdistance02040−2.5−2−1.5−1−0.500.511.5 44 test samples [positives: 6, nonpositives: 38]sample indexdistance(d)By-trajectory.Left:nonconverters;Right:con-verters050−1.5−1−0.500.51 79 test samples [positives: 33, nonpositives: 46]sample indexdistance123456789−1.5−1−0.500.51 9 test samples [positives: 2, nonpositives: 7]sample indexdistance(e)By-CDR.Left:nonconverters;Right:convertersFigure3:Testsetaccuracycomparisonofby-CDRandby-trajectoryclassification.9 (a) Training/test set selection for by-trajectory classification. Sample selection Converters Count Accuracy Count Accuracy 0.86 ± 0.02 By-trajectory Random 45 ± 1 45 ± 1 0.83 ± 0.05 Identical 0.77 ± 0.01 0.85 ± 0.03 (b) Test set Nonconverters 47 ± 2 47 ± 2 Overall accuracy 0.82 ± 0.01 0.84 ± 0.02 Figure 4: By-trajectory average test set classification accuracy for the larger training set size (98 ± 1 per class) and 11, 293 features. and modestly worsened for the "random approach" (from 0.83 ± 0.05 to 0.82 ± 0.01). Evaluating a Third Conversion Definition We now consider a third definition of conversion that combines the first two definitions as follows. Let "converters" consist of individuals who con- verted either by-trajectory or by-CDR (non-white areas in Fig. 2(c)), with the "nonconverter" class consisting of the remaining MCI individuals (white area). The point of view of this new definition, "conversion-by-union", is to be more inclusive in defining an MCI subpopulation at risk, which may ben- efit from early treatment or diagnostic testing. While from that perspective the new definition is reasonable, the fact that grouping individuals by CDR has a role in this definition may be its disadvantage, considering that by-CDR classification was previously shown to perform not much better than random guessing. Results, averaged across the same 10 trials used in Fig. 3, are given in Table 1 and indicate that conversion-by-union generalizes somewhat 20 (a)Training/testsetselectionforby-trajectoryclas-sification.SampleTestsetselectionConvertersNonconvertersOverallCountAccuracyCountAccuracyaccuracyBy-trajectoryRandom45±10.77±0.0147±20.86±0.020.82±0.01Identical45±10.85±0.0347±20.83±0.050.84±0.02(b)Figure4:By-trajectoryaveragetestsetclassificationaccuracyforthelargertrainingsetsize(98±1perclass)and11,293features.−4−2024x 10−300.10.20.30.40.50.60.70.80.91normalized volume of hippocampus(normalized) count 250 [m=−0.015, s=1.004]48 [m=−0.654, s=0.924]155 [m=0.399, s=0.896]143 [m=−0.678, s=0.828]Figure5:Forthehippocampus,by-trajectory(red)haslargerhistogramseparationbetweenconverter(dashedline)andnonconverter(solidline)groupsthanby-CDR(blue).Toillustratethismoreclearly,alsoshownistheGaussiancurveforeachofthesefoursubjectgroups(plottedbasedongroupmean(m)andstandarddeviation(s)indicatedinthefigurelegendwiththesame0.001scalingasthex-axis).10 Sample selection Converters By-union Identical Count Accuracy Count Accuracy 48 ± 1 0.75 ± 0.06 0.78 ± 0.05 Test set Nonconverters 44 ± 2 Overall accuracy 0.77 ± 0.03 Table 1: Average test set accuracy of by-union classification for 39 ± 1 per-class training samples and 11, 293 features. worse than conversion-by-trajectory.17 3.2.2. Validation on Known AD Conversion Biomarkers To validate the proposed conversion definitions with respect to desider- atum 3, we performed correlation tests on the MCI population between the binary class variable C ∈ {0 = no conversion, 1 = conversion} and known AD conversion biomarkers consisting of both 1) volume in reported AD- affected regions (Table 2 in (Schuff et al., 2010)), which we measured for each individual's final-visit MRI18 and 2) the clinical MMSE measure. The stronger the correlation, the more accurately the biomarker is predicted from the class variable and the greater the separation between the biomarker his- tograms, conditioned on the two classes. Before presenting correlation test results, we first illustrate in Fig. 5 the increased separation of the histograms of hippocampus volume for the converter and nonconverter groups in the by-trajectory case, compared with by-CDR. Next, we performed comprehensive statistical tests for a number of suggested AD biomarkers. The R statistical computing package was used to perform all tests with statistical significance set at the 0.05 level. In Table 2a 17Note that "by-union" is, by definition, an instance of the "identical approach", as stated in Table 1. To ensure fairness of comparison with the by-trajectory definition of conversion, our test sets, and training set sizes, in these two cases were identical. In fact, we chose the by-union training set to be as similar to by-trajectory's, in every trial, as possible. Referring to Fig. 3(b) (which represents a trial example), the by-union training set was chosen to include 1) the two large striped groups (red and white); 2) the small "special" gray group (of size seven in this trial example) and its age-matched counterpart within small white-striped group, and; 3) a subset of the second small "special" gray group (two of five individuals in this trial example) and its age-matched counterpart within the small red-striped group. 18As discussed in Appendix A, we measured normalized region volume. Note also that our regions are defined based on the atlas (Atlas2) we used. The correspondence between the regions in (Schuff et al., 2010) and our defined regions is given in Appendix B. Finally, note that a subject's final visit is not always the sixth visit. 21 Figure 5: For the hippocampus, by-trajectory (red) has larger histogram separation be- tween converter (dashed line) and nonconverter (solid line) groups than by-CDR (blue). To illustrate this more clearly, also shown is the Gaussian curve for each of these four subject groups (plotted based on group mean (m) and standard deviation (s) indicated in the figure legend with the same 0.001 scaling as the x-axis). 22 (a)Training/testsetselectionforby-trajectoryclas-sification.SampleTestsetselectionConvertersNonconvertersOverallCountAccuracyCountAccuracyaccuracyBy-trajectoryRandom45±10.77±0.0147±20.86±0.020.82±0.01Identical45±10.85±0.0347±20.83±0.050.84±0.02(b)Figure4:By-trajectoryaveragetestsetclassificationaccuracyforthelargertrainingsetsize(98±1perclass)and11,293features.−4−2024x 10−300.10.20.30.40.50.60.70.80.91normalized volume of hippocampus(normalized) count 250 [m=−0.015, s=1.004]48 [m=−0.654, s=0.924]155 [m=0.399, s=0.896]143 [m=−0.678, s=0.828]Figure5:Forthehippocampus,by-trajectory(red)haslargerhistogramseparationbetweenconverter(dashedline)andnonconverter(solidline)groupsthanby-CDR(blue).Toillustratethismoreclearly,alsoshownistheGaussiancurveforeachofthesefoursubjectgroups(plottedbasedongroupmean(m)andstandarddeviation(s)indicatedinthefigurelegendwiththesame0.001scalingasthex-axis).10 (and Fig. 6), the correlation coefficients for by-trajectory and by-CDR are shown for each biomarker, along with their associated p-values (Chambers et al., 1993). Note that for 11 out of 14 brain regions, the correlation with by-trajectory is greater than the correlation with by-CDR (in bold), with by-trajectory meeting the significance threshold in 10 of these 11 regions. Further, for only two of the remaining four biomarkers - posterior cingulate and the clinical MMSE measure - does the correlation with by-CDR meet the significance threshold. Most notably, well-established markers for AD such as the hippocampus, lateral ventricles, and inferior parietal exhibited strong correlation with the by-trajectory definition. To further assess statistical significance of the comparison between by- trajectory and by-CDR correlations, we performed a correlated correlation test (Meng et al., 1992), the appropriate test given that the same MCI sam- ple population was used in measuring correlations for both by-CDR and by-trajectory. This test (Table 2b) reveals that the larger correlation of by- trajectory is statistically significant at the 0.05 level in nine brain regions (in bold). By contrast, conversion-by-CDR does not achieve a statistically sig- nificant advantage for any of the brain regions, nor with respect to MMSE. To mitigate the confounding effect of age in the AD biomarker measurements, we linearly adjusted them for age prior to performing the experiments de- scribed above. We also repeated these experiments without age adjustment. In this case, the number of regions for which by-trajectory's correlation ex- ceeded by-CDR's rose from 11 to 12 (with the addition of the entorhinal cortex), and the number of regions for which the larger correlation of by- trajectory was statistically significant rose from 9 to 10 (with the addition of the medial orbitofrontal GM). 3.2.3. Identification of prognostic brain "biomarker" regions In the previous section, we validated conversion definitions using estab- lished (diagnostic) AD biomarker brain regions (with volumes measured at final visit). In this section, we will identify key prognostic biomarker brain regions via supervised feature selection, aiming first to identify the "essen- tial" subset of voxel features, i.e. the voxels (at initial visit) necessary for our classifier to well-discriminate the CT and NT classes. The brain regions (consistent with a registered brain atlas) within which these select voxels principally reside then identify our prognostic brain biomarker regions. Sim- ilarly, we will identify diagnostic regions, critical for discriminating between AD and Control subjects (using our AD-Control classifier). In both cases, 23 (a) By-trajectory By-CDR Correlation P-value Correlation P-value Biomarker Entorhinal cortex Fusiform gyrus Hippocampus Inferior parietal GM Lateral orbitofrontal GM Lateral ventricles Medial orbitofrontal GM Parahippocampal gyrus Posterior cingulate Precentral GM Superior frontal GM Superior temporal GM Total GM Total WM MMSE coefficient 0.041 0.430 0.530 0.513 0.077 0.586 0.239 0.095 0.001 0.173 0.331 0.478 0.491 0.170 0.356 0.486 8.20E-15 2.00E-16 2.00E-16 0.185 2.00E-16 3.01E-05 0.102 0.981 0.003 4.50E-09 2.00E-16 2.00E-16 0.003 2.57E-10 coefficient 0.050 0.152 0.231 0.097 0.078 0.211 0.101 0.060 0.144 0.088 0.094 0.161 0.246 0.041 0.453 0.385 0.008 5.67E-05 0.094 0.178 0.000246 0.081 0.307 0.013 0.128 0.104 0.005 1.70E-05 0.477 2.00E-16 (b) P-value 0.889 2.17E-06 2.42E-10 0.985 1.34E-09 0.048 0.616 0.045 0.230 0.001 3.10E-05 0.0001 0.069 1.33E-06 0.118 Table 2: Correlation coefficients and associated p-values: (a) Correlation test results; (b) Correlated correlation test results for each of the regions in (a). Statistically significant results are shown in bold. 24 Figure 6: Comparison of correlation coefficients for by-trajectory and by-CDR. AD-Control classifier only Amygdala left Cingulate region right Entorhinal cortex right Inferior occipital gyrus right Medial occipitotemporal gyrus left Parahippocampal gyrus left Temporal lobe WM right Temporal pole right intersection Hippocampal formation* left Hippocampal formation* right Entorhinal cortex* left Inferior temporal gyrus right Lateral occipitotemporal gyrus* right Parahippocampal gyrus* right Perirhinal cortex left Perirhinal cortex right Middle temporal gyrus right Uncus left CT-NT classifier only Superior temporal gyrus* left Middle temporal gyrus left Precuneus right Lateral front-orbital gyrus* right Insula right Supramarginal gyrus* left Temporal lobe WM left Temporal pole left Medial front-orbital gyrus* left Table 3: Brain regions identified as biomarkers using voxel-based features and MFE. 25 00.10.20.30.40.5correlation coefficient entorhinal cortexhippocampusinferior parietal GMlateral orbitofrontal GMlateral ventriclesmedial orbitofrontal GMparahippocampal gyrusposterior cingulateprecentral GMsuperior frontal GMfusiform gyrustotal GMtotal WMtransverse temporal GMMMSEby−trajectoryby−CDRFigure6:Comparisonofcorrelationcoefficientsforby-trajectoryandby-CDR.Tofurtherassessstatisticalsignificanceofthecomparisonbetweenby-trajectoryandby-CDRcorrelations,weperformedacorrelatedcorrelationtest(Mengetal.,1992),theappropri-atetestgiventhatthesameMCIsamplepopulationwasusedinmeasuringcorrelationsforbothby-CDRandby-trajectory.Thistest(Table2b)revealsthatthelargercorrelationofby-trajectoryisstatisticallysignificantatthe0.05levelinninebrainregions(inbold).Bycontrast,conversion-by-CDRdoesnotachieveastatisticallysignificantadvantageforanyofthebrainregions,norwithrespecttoMMSE.Tomitigatethecon-foundingeffectofageintheADbiomarkermeasurements,welinearlyadjustedthemforagepriortoperformingtheexper-imentsdescribedabove.Wealsorepeatedtheseexperimentswithoutageadjustment.Inthiscase,thenumberofregionsforwhichby-trajectory'scorrelationexceededby-CDR'srosefrom11to12(withtheadditionoftheentorhinalcortex),andthenumberofregionsforwhichthelargercorrelationofby-trajectorywasstatisticallysignificantrosefrom9to10(withtheadditionofthemedialorbitofrontalGM).3.2.3.Identificationofkeyprognosticbrain"biomarker"re-gionsIntheprevioussection,wevalidatedconversiondefini-tionsusingestablished(diagnostic)ADbiomarkerbrainregions(withvolumesmeasuredatfinalvisit).Inthissection,wewillidentifykeyprognosticbiomarkerbrainregionsviasupervisedfeatureselection,aimingfirsttoidentifythe"essential"subsetofvoxelfeatures,i.e.thevoxels(atinitialvisit)necessaryforourclassifiertowell-discriminatetheCTandNTclasses.Thebrainregions(consistentwitharegisteredbrainatlas)withinwhichtheseselectvoxelsprincipallyresidethenidentifyourprognosticbrainbiomarkerregions.Similarly,wewilliden-tifydiagnosticregions,essentialfordiscriminatingbetween(a)By-trajectoryBy-CDRCorrelationP-valueCorrelationP-valueBiomarkercoefficientcoefficientEntorhinalcortex0.0410.4860.0500.385Fusiformgyrus0.4308.20E-150.1520.008Hippocampus0.5302.00E-160.2315.67E-05InferiorparietalGM0.5132.00E-160.0970.094LateralorbitofrontalGM0.0770.1850.0780.178Lateralventricles0.5862.00E-160.2110.000246MedialorbitofrontalGM0.2393.01E-050.1010.081Parahippocampalgyrus0.0950.1020.0600.307Posteriorcingulate0.0010.9810.1440.013PrecentralGM0.1730.0030.0880.128SuperiorfrontalGM0.3314.50E-090.0940.104SuperiortemporalGM0.4782.00E-160.1610.005TotalGM0.4912.00E-160.2461.70E-05TotalWM0.1700.0030.0410.477MMSE0.3562.57E-100.4532.00E-16(b)P-value0.8892.17E-062.42E-100.9851.34E-090.0480.6160.0450.2300.0013.10E-050.00010.0691.33E-060.118Table2:Correlationcoefficientsandassociatedp-values.11 the accuracy of the selected brain region biomarkers rests heavily on the ac- curacy of the supervised feature selection algorithm we employ. In Figures 7(a) and 7(b), we compare MFE and RFE feature elimination (i.e. feature selection via feature elimination) for both Control-AD classification and for CT-NT classification (for one representative, example trial). The curves show test set accuracy as a function of the number of retained features (which is reduced going from right to left). Note that the "MFE/MFE-slack" hybrid method (Aksu et al., 2010)) outperforms RFE for both brain classification tasks, achieving lower test set error rates, and with much fewer retained fea- tures. The circle, determined without use of the test set based on the rule in (Aksu et al., 2010), marks the point at which we stopped eliminating features by MFE, thus determining the (trial's) retained voxel set. This MFE-RFE comparison (and the previous comparison in (Aksu et al., 2010)) supports our use of MFE to determine brain biomarkers. To relate the retained voxel set to anatomic regions in the brain, we overlaid the retained voxel set onto a registered atlas space. For CT-NT classification, to improve robustness, the final voxel set was formed from the union of the retained voxel sets from each of ten feature elimination trials (each using a different, randomly selected training sample subset). For AD-Control classification, the final voxel set came from a single trial (the only trial, from which the 10 CT-NT trials stemmed). For each of these two cases, overlaying the final voxel set onto the co-registered atlas (Atlas2, defined in Appendix A) yielded between 70-80 anatomic regions. For data interpretation purposes, we then identified a subset of (biomarker) regions using the following procedure. First, for each brain region, we measured the percentage of the region's voxels that are retained, sorted these percentages, and then plotted them. As shown in Fig. 8(a), the resulting curve for the AD-Control case has a distinct knee, which we thus used as a threshold (0.125) to select the final, retained (diagnostic) regions for AD-Control. We used the same threshold for the CT-NT curve, shown in Fig. 8(b). This choice of threshold yields a reasonable number of regions -- 19 for the CT-NT (prognostic) case and 21 for the AD-Control (diagnostic) case. The resulting sets of identified prognostic and diagnostic biomarkers are given in Table 3, along with their intersection. The diagnostic markers in the table include the majority of the known brain regions in the medial tempo- ral lobe involved in AD pathology. For example, hippocampus atrophy and lateral ventricle enlargement, particularly in its anterior aspects of the tem- poral horn, are considered the most prominent diagnostic markers for AD. 26 Entorhinal cortical regions, including the perirhinal cortex, are presumably the earliest sites of degeneration (Braak et al., 1997). Thus, independent identification by our AD-control classifier of known AD diagnostic biomark- ers establishes a reasonable basis for applying the same approach to identify prognostic biomarkers. The brain regions listed as CT-NT prognostic mark- ers include most known AD diagnostic markers (including 8 of the 12 regions from (Schuff et al., 2010) (marked by *), 4 of which are also diagnostic mark- ers), indicating that some AD-linked pathological changes in these brain re- gions already occurred and remained active in a subset of MCI subjects who likely progress to AD rapidly. Conversely, the brain areas appearing only on the prognostic marker list are likely the most active areas of degeneration during this stage of progression to dementia. These structures tend to be the brain regions further away from the entorhinal cortex onto the parietal (Supramarginal gyrus, Precuneus) and temporal cortex (Superior temporal gyrus and Middle temporal gyrus) regions. All the brain structures listed in the table are known to be involved in AD (Braak et al., 1997; Chan et al. , 2003; Frisoni et al., 2009). Thus, the markers in Table 3 suggest an interesting anatomic pattern of trajectory for MCI conversion to AD which conforms with the Brak and Brak hypothesis and previous imaging findings (Chan et al. , 2003; Frisoni et al., 2009). Moreover, the CT-NT regions uniquely found by our MFE-based procedure in Table 3 may be viewed as "putative" prognostic markers, and may warrant further investigation. Finally, we note that we have used a particular criterion (percentage of a region's voxels that are retained) to identify biomarker regions, starting from MFE-retained voxels. While our identified regions are plausible, it is possible that other (equally reasonable) criteria may produce different biomarker re- gion results. Thus, the biomarkers we identify should be viewed as anecdotal, identifying regions that figure prominently in our classifier's decisionmaking and also potentially assisting researchers in forming hypotheses about MCI- to-AD disease progression. However, we do not view the identified regions as definitive. 3.2.4. Comparison with an SPM-based biomarker identification approach In the previous section, we used MFE to identify voxels as biomarkers for the CT and NT classes. Here, using the same CT-NT training and test populations, we will alternatively identify voxel-based biomarkers us- ing statistical testing with SPM5 (see: (SPM website)). Subsequently we will present a classifier generalization accuracy comparison (where accuracy 27 (a) (b) Figure 7: Test set misclassification rate during the course of feature elimination, for (a) the AD-Control classifier and (b) the CT-NT classifier. (a) (b) Figure 8: Sorted retained voxel percentages for initial regions (a) AD-Control; b) CT-NT, used to select final regions (Sec. 3.2.3). 28 02000400060008000100001200000.10.20.30.40.50.60.70.8number of features retained (starting at 11293)test set classification error rate RFEMFE/MFE(cid:239)slack02000400060008000100001200000.10.20.30.40.5number of features retained (starting at 11293)test set classification error rate RFEMFE/MFE(cid:239)slack(a)(b)Figure7:[ABOVEARENOTTHEACTUALIMAGES]Imagedisplayofbrainregionsidentifiedasbiomarkersusingvoxel-basedfeaturesandMFE.02000400060008000100001200000.10.20.30.40.50.60.70.8number of features (starting at 11293)test set classification error rate RFEMFE/MFE−slack(a)02000400060008000100001200000.10.20.30.40.5number of features (starting at 11293)test set classification error rate RFEMFE/MFE−slack(b)Figure8:Testsetmisclassificationrateduringthecourseoffeatureelimination,for(a)theAD-Controlclassifierand(b)theCT-NTclassifier.02040608010000.10.20.30.40.50.60.70.8ratioindex(a)0102030405060708000.050.10.150.20.250.30.350.40.450.5ratioindex(b)Figure9:Sortedretainedvoxelpercentagesforinitialregions(a)AD-Control;b)CT-NT,usedtoselectfinalregions(Sec.3.2.3).13(a)(b)Figure7:[ABOVEARENOTTHEACTUALIMAGES]Imagedisplayofbrainregionsidentifiedasbiomarkersusingvoxel-basedfeaturesandMFE.02000400060008000100001200000.10.20.30.40.50.60.70.8number of features (starting at 11293)test set classification error rate RFEMFE/MFE−slack(a)02000400060008000100001200000.10.20.30.40.5number of features (starting at 11293)test set classification error rate RFEMFE/MFE−slack(b)Figure8:Testsetmisclassificationrateduringthecourseoffeatureelimination,for(a)theAD-Controlclassifierand(b)theCT-NTclassifier.02040608010000.10.20.30.40.50.60.70.8ratioindex(a)0102030405060708000.050.10.150.20.250.30.350.40.450.5ratioindex(b)Figure9:Sortedretainedvoxelpercentagesforinitialregions(a)AD-Control;b)CT-NT,usedtoselectfinalregions(Sec.3.2.3).13 is again measured on the previous section's CT-NT (test) population) for these two biomarker detection methods. We determined SPM biomarkers as follows. The CT-NT training set population, being age-matched, is read- ily suitable for a paired t-test, an appropriate statistical test for determining SPM-identified biomarkers, i.e. voxels that discriminate between the CT and NT groups. In contrast with MFE's use of only one (out of 216) RAVENS subsamples (taken jointly from the GM, WM, ventricle maps), we performed t-tests on whole RAVENS maps (without subsampling), which makes the SPM-MFE comparison favorably biased towards SPM. More specifically our steps were as follows. First, for the GM and WM maps separately, we found using SPM that a large portion of each of these two tissues was statistically significant at the 0.05 level when correction for multiple comparisons was not applied. Next, we used SPM's FDR-based correction for multiple com- parisons -- based on an SPM FDR cluster size of 5 voxels we found that the spatial extent of the statistically significant regions, at each of the levels 0.05, 0.01, and 0.005, was approximately a subset of the above-mentioned spatial support found in the uncorrected case. Given that the number of significant voxels in any of these SPM experiments is, again due to no subsampling, much larger than the 11,293 voxels started from in the MFE case, we sim- ply 1) chose as our SPM result the result for 0.01 (FDR-corrected), 2) took from among those significant voxels the most significant 11,434 voxels in or- der to be able to compare MFE and SPM for ≈ the same number of voxels (biomarkers). To obtain the generalization accuracy for this SPM-identified biomarker voxel set, using the same training/test set as in the MFE experi- ment, we trained an SVM classifier and measured its generalization accuracy, which was found to be 0.76. This accuracy is somewhat lower than the 0.8 accuracy of the previous section's CT-NT SVM classifier. Recalling that this comparison is actually favorably biased towards SPM, and further noticing the fact that MFE was able to maintain the 0.8 accuracy all the way down to 2000 features (cf. Fig. 7(b)), this experimental comparison provides another validation (beyond the comparison with RFE given earlier) for MFE-based feature/biomarker selection, applied to brain images. 4. Conclusions We have presented an automated prognosticator of MCI-to-AD conver- sion based on brain morphometry derived from high resolution ADNI MR images. The primary novel contributions of our work are: i) casting MCI 29 prognostication as a novel machine learning problem lying somewhere be- tween supervised and unsupervised learning; ii) our proposal of a conversion definition which, unlike previous methods, exploits both rich phenotypical information in neuroimages and AD and control examples; iii) correlation testing and classifier accuracy evaluations to validate candidate conversion definitions; iv) prognostic biomarker discovery based on our conversion defi- nition. We demonstrated that our method achieved both better generaliza- tion accuracy and stronger, statistically significant, correlations with known brain region biomarkers than a predictor based on the clinical CDR score, the approach used in several past works. The brain structures identified as AD-control diagnostic markers and MCI conversion prognostic markers well conform with known brain atrophic patterns and progression trajecto- ries occurring in AD-afflicted brains. While the noisy nature of cognitive assessments, including MMSE, has been acknowledged in past works, in fu- ture we may extend our methodology to consider cognitive assessment data, both potentially as additional (baseline) input features and as additional or alternative prediction targets to our "conversion-by-trajectory" labels. We may also consider alternative ways to adjust for confounding effects of age, noting that (Schuff et al., 2010) has characterized the nonlinear dependence of age on brain region volumes. Finally, while we have focused on the MCI subpopulation here, our system could also potentially be used to detect, as possible misdiagnoses, subjects diagnosed as "Control" who are classified as MCI converters by our system. 5. Acknowledgement Funding supporting this work has been provided in part through NIH R01 AG02771 and the Pennsylvania Department of Health. We thank Dr. Michelle Shaffer for assistance with statistical testing and Jianli Wang and Zachary Herse for assistance with segmentation. Data collection and sharing for this project was funded by the Alzheimer's Disease Neuroimaging Initiative (ADNI) (National Institutes of Health Grant U01 AG024904). ADNI is funded by the National Institute on Aging, the National Institute of Biomedical Imaging and Bioengineering, and through generous contributions from the following: Abbott, AstraZeneca AB, Bayer Schering Pharma AG, Bristol-Myers Squibb, Eisai Global Clinical Devel- opment, Elan Corporation, Genentech, GE Healthcare, GlaxoSmithKline, Innogenetics, Johnson and Johnson, Eli Lilly and Co., Medpace, Inc., Merck 30 and Co., Inc., Novartis AG, Pfizer Inc, F. Hoffman-La Roche, Schering- Plough, Synarc, Inc., as well as non-profit partners the Alzheimer's Asso- ciation and Alzheimer's Drug Discovery Foundation, with participation from the U.S. Food and Drug Administration. Private sector contributions to ADNI are facilitated by the Foundation for the National Institutes of Health (www.fnih.org). The grantee organization is the Northern California In- stitute for Research and Education, and the study is coordinated by the Alzheimer's Disease Cooperative Study at the University of California, San Diego. ADNI data are disseminated by the Laboratory for Neuro Imaging at the University of California, Los Angeles. This research was also supported by NIH grants P30 AG010129, K01 AG030514, and the Dana Foundation. Appendix A. Image processing The input to our processing is a T1-weighted MR image of the head. First, using rigid-body registration implemented in FSL's linear image reg- istration tool FLIRT (Jenkinson et al., 2001), we coarsely aligned this (3d) image with the MNI/ICBM atlas resampled to 1mm isotropic voxel dimen- sions (aka Atlas1). We used FSL version 3 (see: (FSL website)). Next, we removed non-brain anatomy from the aligned head image, using FSL's brain extraction tool BET (Smith et al., 2002). We then segmented the resulting brain-only 3d image into the following five segments required by HAMMER(Shen et al., 2002)19 : WM, GM, ventricles, (non-ventricle) CSF, (non-anatomy) background. Next, using as inputs an MNI atlas distributed with HAMMER (aka Atlas220) and the ADNI participant's five-segment im- age, we performed two different HAMMER operations, generating: 1) the three "volumetric density" Ravens images: GM, WM, ventricles. The union of these three images (following some preprocessing) forms the set of features used by our classifier; 2) the 3d region-segmented image, whose region vol- umes are used for by-trajectory statistical validation in section 3.2.2 and for region biomarker identification in section 3.2.3.21 19We used a 2006 version of HAMMER, downloaded on November 8, 2006. 20Since the region-segmented version of the atlas distributed with HAMMER had much better segmentation quality than the five-segment version, as a pre-processing step we used the former to create a replacement for the latter (by simply taking the union of all the regions). The resulting five-segment atlas is dubbed 'Atlas2'. 21We measured normalized region volumes. We normalized by dividing by the sum of all (98) region volumes. This sum is essentially intracranial volume minus cerebellum volume, 31 Each of these two HAMMER operations performs a different type of atlas- based nonlinear registration -- the one that generates the RAVENS tissue maps performs less aggressive warping between the five-segment image and the atlas in order both to combat noise inherent in the registration process and to preserve volume on a tissue-by-tissue basis (so as to properly detect and encode brain atrophy). Consequently, there is considerable individual variability in RAVENS images, which we mitigate in a standard way by smoothing with a Gaussian filter with an FWHM of 5mm. The sum of the voxel intensities over the three RAVENS maps is on the order of 106 and varies across individuals. Thus, prior to smoothing, we normalized each individual's Ravens maps, such that each individual has the same total volume. However, after the normalization and smoothing (aka "nsRAVENS" images), some areas of poor registration, manifesting as areas with very low voxel values (including zero), will remain.22 For a population of nsRAVENS images, these areas are considered to be outlier areas and are thus removed from each image in the population by thresholding 23. The resulting images are the features used by our classification system. Appendix B. Correspondence Between Atlas-defined Regions and Those Defined in (Schuff et al., 2010) as our "intracranial region" list includes CSF in addition to brain regions and excludes the cerebellum. 22A substantial portion of the very low, nonzero voxel intensities are in fact introduced by the smoothing itself as it calculates each new voxel intensity as a weighted average of many neighboring voxels (thereby switching some voxels from zero (non-anatomy) to very low intensity values (anatomy), which essentially slightly grows the anatomy boundaries outward). 23We calculated the threshold solely using the training population of our AD-Control classifier, and then applied the thresholding operation on the entire population of AD, MCI, and Control individuals. In this way, we were careful to exclude test examples from all phases of classifier training. 32 Entorhinal cortex left/right Lateral occipitotemporal gyrus right/left Hippocampal formation right/left Supramarginal gyrus left/right, Angular gyrus right/left Lateral ventricle left/right Entorhinal cortex Fusiform gyrus Hippocampus Inferior parietal GM Lateral orbitofrontal GM Lateral front-orbital gyrus right/left Lateral ventricles Medial orbitofrontal GM Medial front-orbital gyrus right/left Parahippocampal gyrus Posterior cingulate Precentral GM Superior frontal GM Superior temporal GM Parahippocampal gyrus right/left Cingulate region left/right Precentral gyrus right/left Superior frontal gyrus left/right Superior temporal gyrus right/left Table B.4: Correspondence between the regions in (Schuff et al., 2010) (left) (except "Total GM" and "Total WM") and our defined regions (right). References Y. Aksu, D. J. Miller, G. Kesidis, and Q. X .Yang, "Margin-Maximizing Feature Elimination Methods for Linear and Nonlinear Kernel-Based Dis- criminant Functions", IEEE Transactions on Neural Networks, vol. 25, no.10, pp.701-717, 2010. H. Braak and E. Braak, "Frequency of Stages of Alzheimer-Related Lesions in Different Age Categories", Neurobiology of Aging 18(4):351-357, July 8 1997. C. Chang and C. Lin, "LIBSVM : a library for support vector machines," software available at http://www.csie.ntu.edu.tw/∼cjlin/libsvm, 2001. J. M. Chambers, "Linear models", in: Statistical Models in S, J. M. Chambers and T. J. Hastie (Eds.), Chapman & Hall, New York, 1993. D. Chan, J.C. Janssen, J.L. Whitwell, H.C. Watt, R. Jenkins, C. Frost, M.N. Rossor and N.C. Fox, "Change in rates of cerebral atrophy over time in early-onset Alzheimer's disease: longitudinal MRI study", Lancet 362, pp. 1121-1122, 2003. G. Chetelat, B. Desgranges, V. de la Sayette, F. Viader, F. Eustache, J-C. Baron, "Mapping gray matter loss with voxel-based morphometry in mild cognitive impairment", NeuroReport vol.13. no.15, 28 October 2002. Y.-Y. Chou, N. Lepor´e, C. Avedissian, S. K. Madsen, X. Hua, C. R. Jack Jr., M. W. Weiner, A. W. Toga, P. M. Thompson, and the Alzheimer's 33 Disease Neuroimaging Initiative, "Mapping Ventricular Expansion and its Clinical Correlates in Alzheimer's Disease and Mild Cognitive Impairment using Multi-Atlas Fluid Image Alignment", Proc. SPIE, vol.7259, 725930, 2009. Y.-Y. Chou, N. Lepor´e, P. Saharan, S. K. Madsen, X. Hua, C. R. Jack, L. M. Shaw, J. Q. Trojanowski, M. W. Weiner, A. W. Toga, P. M. Thompson, and the Alzheimer's Disease Neuroimaging Initiative, "Ventricular maps in 804 ADNI subjects: correlations with CSF biomarkers and clinical de- cline", Neurobiology of Aging 31, pp.1386-1400, 2010. J. G. Csernansky, L. Wang, J. Swank, J. P. Miller, M. Gado, D. McKeel, M. I. Miller, J. C. Morris, "Preclinical detection of Alzheimer's disease: hippocampal shape and volume predict dementia onset in the elderly", NeuroImage 25, pp.783-792, 2005. C. Davatzikos, "Mapping image data to stereotaxic spaces: Applications to brain mapping," Hum. Brain Mapp., 6:334-338, 1998. C. Davatzikos, A. Genc, D. Xu, and S. M. Resnick, "Voxel-Based Morphom- etry Using the RAVENS Maps: Methods and Validation Using Simulated Longitudinal Atrophy," NeuroImage 14, pp.1361-1369, 2001. C. Davatzikos, Y. Fan, X. Wu, D. Shen, S. M. Resnick, "Detection of pro- dromal Alzheimer's disease via pattern classification of magnetic resonance imaging", Neurobiology of Aging 29, pp.514-523, 2008. C. Davatzikos, P. Bhatt, L. M. Shaw, K. N. Batmanghelich, J. Q. Trojanowski, "Prediction of MCI to AD conversion, via MRI, CSF biomarkers, and pattern classification", Neurobiology of Aging, 2010, doi:10.1016/j.neurobiolaging.2010.05.023. M. J. de Leon, S. DeSanti, R. Zinkowski, P. D. Mehta, D. Pratico, S. Segal, H. Rusinek, J. Li, W. Tsui, L. A. Saint Louis, C. M. Clark, C. Tarshish, Y. Li, L. Lair, E. Javier, K. Rich, P. Lesbre, L. Mosconi, B. Reisberg, M. Sadowski, J. F. DeBernadis, D. J. Kerkman, H. Hampel, L. -O. Wahlund, P. Davies, "Longitudinal CSF and MRI biomarkers improve the diagnosis of mild cognitive impairment," Neurobiology of Aging 27, pp.394-401, 2006. 34 G. De Meyer, F. Shapiro, H. Vanderstichele, E. Vanmechelen, S. Engel- borghs, P. P. De Deyn, E. Coart, O. Hansson, L. Minthon, H. Zetter- berg, K. Blennow, L. Shaw, J. Q. Trojanowski, for the Alzheimers Dis- ease Neuroimaging Initiative, "Diagnosis-Independent Alzheimer Disease Biomarker Signature in Cognitively Normal Elderly People", Arch. Neu- rol., vol. 67, no. 8, pp.949-956, Aug 2010. R. Duda, P. Hart, and G. Stork, Pattern Classification. Second Edition, John Wiley and Sons, New York, 2001. Y. Fan, D. Shen, R. C. Gur, R. E. Gur, C. Davatzikos, "COMPARE: Clas- sification of Morphological Patterns Using Adaptive Regional Elements", IEEE Transactions on Medical Imaging, vol. 26, no. 1, pp.93-105, 2007. Y. Fan, N. Batmanghelich, C. M. Clark, C. Davatzikos, "Spatial patterns of brain atrophy in MCI patients, identified via high-dimensional pat- tern classification, predict subsequent cognitive decline", NeuroImage 39, pp.1731-1743, 2008. C. Fennema-Notestine, D. J. Hagler Jr., L. K. McEvoy, A. S. Fleischer, E. H. Wu, D. S. Karow, A. M. Dale, the Alzheimer's Disease Neuroimaging Initiative, "Structural MRI biomarkers for preclinical and mild Alzheimer's disease", Human Brain Mapping, vol.30, issue 10, pp.3238-3253, 2009. G.B. Frisoni, A. Prestia, P.E. Rasser, M. Bonetti, P.M. Thompson, "In vivo mapping of incremental cortical atrophy from incipient to overt Alzheimer's disease", Neurol. 256(6):916-24, Feb 28, 2009. FSL website: FSL http://www.fmrib.ox.ac.uk/fsl (FMRIB Software Library), G. Fung, O. L. Mangasarian, "A feature selection Newton method for support vector machine classification," Computational Optimization and Applica- tions, vol. 28, no.2:185-202, July 2004. A. F. Goldszal, C. Davatzikos, D. L. Pham, M. X. H. Yan, R. N. Bryan, S. M. Resnick, "An Image-Processing System for Qualitative and Quantitative Volumetric Analysis of Brain Images," J. Comput. Assist. Tomogr., 22:827- 837, 1998. 35 I. Guyon, J. Weston, S. Barnhill, and V. Vapnik, "Gene selection for cancer classification using support vector machines," Machine Learning, 46(1):389-422, 2002. I. Guyon and A. Elisseeff, "An introduction to variable and feature selection," J. Mach. Learn. Res., vol. 3, pp.1157-1182, 2003. M. Jenkinson and S.M. Smith, "A global optimisation method for robust affine registration of brain images," Med. Image Anal., 5(2):143-156, 2001. J. P. Lerch, A. C. Evans, "Cortical thickness analysis examined through power analysis and a population simulation", NeuroImage 24, pp.163-173, 2005. A. D. Leow, I. Yanovsky, N. Parikshak, X. Hua, S. Lee, A. W. Toga, C. R. Jack Jr., M. A. Bernstein, P. J. Britson, J. L. Gunter, C. P. Ward, B. Borowski, L. M. Shaw, J. Q. Trojanowski, A. S. Fleisher, D. Harvey, J. Kornak, N. Schuff, G. E. Alexander, M. W. Weiner, P. M. Thompson, Alzheimers Disease Neuroimaging Initiative. "Alzheimers Disease Neu- roimaging Initiative: A One-Year Follow-up Study Using Tensor-Based Morphometry Correlating Degenerative Rates, Biomarkers and Cogni- tion," Neuroimage 45, pp.645-55, 2009. X. -L. Meng, R. Rosenthal, D. B. Rubin, "Comparing Correlated Correlation Coefficients", Psychological Bulletin 111, pp.172-175, 1992. C. Misra, Y. Fan, C. Davatzikos, "Baseline and longitudinal patterns of brain atrophy in MCI patients, and their use in prediction of short-term conver- sion to AD: Results from ADNI", NeuroImage 44, pp.1415-1422, 2009. R. C. Petersen, Mild cognitive impairment: aging to Alzheimer's Disease, Oxford University Press, 2004. H. Rusinek, Y. Endo, S. De Santi, D. Frid, W.-H. Tsui, S. Segal, A. Convit, "Atrophy rate in medial temporal lobe during progression of Alzheimer's disease", Neurology, vol.63, issue 12, 2354-2359, 2004. J. M. Schott, S. L. Price, C. Frost, J. L. Whitwell, et al, "Measuring atro- phy on Alzheimer disease - A serial MRI study over 6 and 12 months," Neurology, 65(1), pp.119-124, 2005. 36 N. Schuff, D. Tosun, P. S. Insel, G. C. Chiang, D. Truran, P. S. Aisen, C. R. Jack, Jr., M. W. Weiner, the Alzheimer's Disease Initiative, "Nonlinear time course of brain volume loss in cognitively normal and impaired elders", Neurobiology of Aging, 2010, doi:10.1016/j.neurobiolaging.2010.07.012. S. M. Smith, "Fast robust automated brain extraction," Human Brain Map- ping, 17(3):143-155, 2002. D. Shen and C. Davatzikos, "HAMMER: hierarchical attribute match- ing mechanism for elastic registration," IEEE Trans. Medical Imaging, 21(11):1421-1439, 2002. D. Shen, C. Davatzikos, "Very High-resolution Morphometry Using Mass- preserving Deformations and HAMMER Elastic Registration," NeuroIm- age 18, pp.28-41, 2003. SPM website: Statistical http://www.fil.ion.ucl.ac.uk/spm Parametric Mapping (SPM), T. R. Stoub, M. Bulgakova, S. Leurgans, D. A. Bennett, D. Fleischman, D. A. Turner, L. deToledo-Morrell, "MRI predictors of risk of incident Alzheimer disease," Neurology 64, pp. 1520-1524, 2005. P. M. Thompson, K. M. Hayashi, G. de Zubicaray, A. L. Janke, S. E. Rose, et al, "Dynamics of gray matter loss in Alzheimer's disease," J. Neurosci. 23, pp.994-1005, 2003. V. Vapnik, Statistical Learning Theory. John Wiley & Sons, 1998. P. Vemuri, J. L. Gunter, M. L. Senjem, J. L. Whitwell, K. Kantarci, D. S. Knopman, B. F. Boeve, R. C. Petersen, C. R. Jack Jr., "Alzheimer's disease diagnosis in individual subjects using structural MR images: Validation studies," NeuroImage 39, pp.1186-1197, 2008. P. Vemuri, H. J. Wiste, S. D. Weigand, L. M. Shaw, J. Q. Trojanowski, M. W. Weiner, D. S. Knopman, R. C. Petersen, C. R. Jack Jr., and On behalf of the Alzheimer's Disease Neuroimaging Initiative, "MRI and CSF biomarkers in normal, MCI, and AD subjects: Predicting future clinical change," Neurology 73:294-301, 2009. 37 Y. Wang, Y. Fan, P. Bhatt, C. Davatzikos, "High-dimensional pattern regres- sion using machine learning: From medical images to continuous clinical variables," NeuroImage 50, pp.1519-1535, 2010. Y. Zhang, M. Brady, and S. Smith, "Segmentation of brain MR images through a hidden Markov random field model and the expectation maxi- mization algorithm," IEEE Trans. Medical Imaging, 20(1):45-57, 2001. D. Zhang, Y. Wang, L. Zhou, H. Yuan, D. Shen, and the Alzheimer's Disease Neuroimaging Initiative, "Multimodal classification of Alzheimer's disease and mild cognitive impairment," NeuroImage, 2011, doi:10.1016/j.neuroimage.2011.01.008. 38
1208.5350
4
1208
2013-02-19T17:08:00
Impact of intrinsic biophysical diversity on the activity of spiking neurons
[ "q-bio.NC" ]
We study the effect of intrinsic heterogeneity on the activity of a population of leaky integrate-and-fire neurons. By rescaling the dynamical equation, we derive mathematical relations between multiple neuronal parameters and a fluctuating input noise. To this end, common input to heterogeneous neurons is conceived as an identical noise with neuron-specific mean and variance. As a consequence, the neuronal output rates can differ considerably, and their relative spike timing becomes desynchronized. This theory can quantitatively explain some recent experimental findings.
q-bio.NC
q-bio
Impact of intrinsic biophysical diversity on the activity of spiking neurons Bernstein Center Freiburg & Faculty of Biology, University of Freiburg, Freiburg, Germany Man Yi Yim, Ad Aertsen, Stefan Rotter∗ (Dated: May 7, 2018) We study the effect of intrinsic heterogeneity on the activity of a population of leaky integrate- and-fire neurons. By rescaling the dynamical equation, we derive mathematical relations between multiple neuronal parameters and a fluctuating input noise. To this end, common input to het- erogeneous neurons is conceived as an identical noise with neuron-specific mean and variance. As a consequence, the neuronal output rates can differ considerably, and their relative spike timing becomes desynchronized. This theory can quantitatively explain some recent experimental findings. I. INTRODUCTION In statistical physics, it is often assumed that indi- viduals are intrinsically identical. In neuroscience also, identical parameters are typically assumed for all neu- rons in the study of neuronal population activity and correlation transmission. Real neurons, though, even if they are of the same type and located in the same brain area, exhibit intrinsic differences. Their morphologies and the intracellular concentrations of ions, to name just two examples, can differ widely, although in principle they have been generated by the same mechanisms [1]. As a consequence, neuronal spike patterns can differ al- though neurons receive identical inputs [2, 3]. Recently, in vitro intracellular recordings of isolated mitral cells in the mouse olfactory bulb were conducted while they re- sponded to identical input [3] (Fig. 1(a)). The neurons displayed diverse output firing rates and pairwise corre- lations. Specifically, the spike correlation coefficient ob- tained with a 1 ms observation window co-varied with the rate difference of the neuron pairs: small differences re- sulted in a wide range of different spike correlations, but large differences led always to small spike correlation. In homogeneous network models, additional indepen- dent Gaussian white noises or independent Poisson spikes are very often added to every constituent identical neu- ron to account for their diverse spike timing. In real brain networks, not only the spike timing but also the spiking rate of neurons differ due to their intrinsic bio- physical diversity. Therefore, it is of great interest to understand how the biophysical heterogeneity of a neu- ronal population contributes to neural coding and infor- mation processing in neuronal networks. Research work has been conducted on the coding properties [4, 5] and synchronous responses [6 -- 8] in a network of heteroge- neous neurons. In many cases, neuronal heterogeneity was implemented simply by replacing one or more fixed neuronal parameters, such as the offset current [6, 7], the spiking threshold [5], or the synaptic conductance [4], by a Gaussian- or uniformly-distributed random variable. Here we investigated more fundamental questions, us- ing both theoretical analysis and simulations: how neu- ∗ [email protected] ronal heterogeneity can be represented appropriately in theory and how it can affect the neuronal dynamics and the spiking statistics in a population of simple leaky integrate-and-fire (LIF) neurons. The limitations of the existing approaches are addressed first. Then we suggest a more general scheme to implement biophysical diver- sity when either rate or correlation is of interest. By rescaling the dynamical equation, we derive mathemati- cal relations between multiple neuronal parameters and the input noise. The main impact of common input to heterogeneous neurons on rate and correlation can be realized by an identical (frozen) noise current injection with different values of mean and variance, whereas the complete effect is captured by additionally drawing dis- tributed values of the membrane time constant and the refractory period. In this scheme, the rate difference of heterogeneous LIF neurons can be treated analytically. As for correlation, we utilize alternative correlation mea- sures to illustrate that spikes from heterogeneous neurons may be desynchronized by several milliseconds, thus es- caping detection by a 1 ms observation window. II. MODEL We consider a population of isolated leaky integrate- and-fire (LIF) neurons, each of which has its membrane potential V (t) governed by τm V (t) = −V (t) + RI(t), (1) where the input synaptic current RI(t) = τmJE Xj δ(t − tj) − τmJI Xk δ(t − tk). (2) τm = RC is the membrane time constant. R and C are the membrane resistance and capacitance, respectively. JE (JI ) is the amplitude of an excitatory (inhibitory) post-synaptic potential, whereas tj (tk) represents the time of the jth (kth) excitatory (inhibitory) input spike. When V (t) = θ, V (t) is reset to Vr and a pause for synap- tic integration τr is imposed to mimic the refractory pe- riod. In the high-input regime, the sum of synaptic in- puts to a neuron can be approximated by a fluctuating input noise [9, 10] I(t) ≡ τm[µ + ση(t)], (3) a      b 1.0 0.5 r r o C e k p S i 0.0 0 40% 20% 10% 20 60 Rate difference (Hz) 40 80 FIG. 1. (Color online). (a) Correlation coefficient (1 ms win- dow) of 589 spike trains from mitral cells in vitro receiving an identical input as a function of rate difference, adapted with permission from [3]. (b) The same measure for 100 simulated heterogeneous LIF neurons using uniformly distributed τm, τr, C and θ with different distribution widths. where µ = JEνE − JI νI , σ = qJ 2 EνE + J 2 I νI . (4) (5) is a white noise random process η(t) such that hη(t)η(t′)i = δ(t − t′). νE (νI ) is the firing rate of the excitatory (inhibitory) input. integration of Eq. 1 in our simula- tions was performed using the fourth-order Runge-Kutta method with a time step of 0.01 ms. The numerical III. HETEROGENEITY Intrinsic diversity of a population of neurons can be directly imposed by drawing neuronal parameters from a distribution. Here, we tested the response of 100 isolated heterogeneous LIF neurons to an identical fluctuating in- put current in the form of Eq. (3). Neuronal heterogene- ity is implemented by drawing four uniformly distributed parameters: τm, τr, C (or R) and θ, which, together with Vr, represent all the independent parameters of a LIF neuron in response to a current input. The mean values of the uniform distribution are 20 ms, 2 ms, 1 and 1 (in arbitrary units) respectively, and Vr is fixed to 0. These values are used throughout this work. We maintain the temporal scale of the dynamics of a typical neuron and rescale the potential by setting the mean reset to zero and the mean threshold to 1. The correlation coefficient as a function of the output firing rate difference of all possible pairs with different distribution widths (percentage with respect to the mean) from 100 s of simulations (an exam- ple with µ = 0.03 and σ = 0.3 shown in Fig. 1(b)) highly resemble the experimental findings in [3] (Fig. 1(a)). Diverse neuronal spike timing in a network has very of- ten been achieved by adding independent random inputs to individual neurons. We provide every identical neuron with a common input as the input signal plus an indepen- dent input with the same statistics among neurons , in a 100 D I n o r u e N 50 0 30 t n u o C 0 100 120 Time (ms) 140 1.0 0.5 r r o C e k p S i 0.0 0 2 b 100 50 0 30 0 1.0 0.5 100 120 Time (ms) 140 c 100 50 0 30 0 1.0 0.5 100 120 Time (ms) 140 60 20 Rate difference (Hz) 40 0.0 0 80 20 60 Rate difference (Hz) 40 0.0 0 80 20 60 Rate difference (Hz) 40 80 FIG. 2. Rasterplot, peri-stimulus time histogram (PSTH) and correlation coefficient as a function of rate difference for the following input into 100 LIF neurons: (a) µ + σ(cid:2)√cη(t) + √1 − cξi(t)(cid:3) (c = 0.9); (b) identical input µ + ση(t) but dis- tributed θi; (c) our scheme using µi + σiη(t) together with distributed τm and τr, consistent with both the experimental findings [3] and the mathematical analysis. The subscript i denotes "independent". the form of µ + σ(cid:2)√cη(t) + √1 − cξi(t)(cid:3) where η(t) and ξi(t) are independent Gaussian white noises. Fig. 2(a) displays the raster and the correlation coefficient as func- tions of the rate difference when µ = 0.06, σ = 0.2 and c = 0.9. The rate difference is close to zero and the cor- relation coefficient between any pair is nearly the same [11]. Decreasing c leads to a drop in spike correlation but has no effect on the rate difference. These observations are very distinct from both the experimental (Fig. 1(a)) and simulation (Fig. 1(b)) results. In view of some previ- ous work on the reliability of single neurons in response to a repeated input [2, 3, 12], this implementation may be adopted to account for trial-to-trial variability of the same neuron. It is common practice to implement heterogeneity of neurons by drawing random variables for a single neu- ronal parameter. To test its validity, we provide every identical neuron with a common input plus a random value of the spiking threshold drawn from a uniform dis- tribution θ ∈ [0.5, 1.5]. Fig. 2(b) displays the raster and the correlation coefficient as functions of the rate differ- ence. Another example with distributed values of the in- put offset current µ instead of θ is shown in Fig. 3(a). In either case, neurons exhibit dispersive firing rates. How- ever, the spike correlation distribution at different values of rate difference is too narrow compared with Fig. 1(a) and (b), and the region of small rate difference and small spike correlation cannot be reached. Thus, the impact of neuronal heterogeneity is only partially accounted for. We describe our alternative approach in the next section. IV. MATHEMATICAL RELATIONS In addition to the five parameters mentioned above, two additional parameters correspond to synaptic inputs: JE and JI . We analyze the contribution of the hetero- geneity in the seven independent neuronal parameters in a population of LIF neurons in the high-input regime. Regarding synaptic inputs, heterogeneity in the pa- rameters JE and JI can be captured by heterogeneity in µ and σ, according to Eqs. (4) and (5). For exam- ple, if JE is uniformly distributed, µ is also uniformly distributed, whereas σ2 is distributed as the square of a uniformly distributed variable. The other five parameters are present in the neuronal dynamics irrespective of the type of inputs (current or spikes). First, R (or C, depending on the form of writing) can be absorbed into I(t) as shown in Eq. (1) so any distribution of R can be accounted for by a corresponding distribution of µ and σ. The difference between θ and Vr, which is the potential difference a neuron has to traverse, is a quantity relative to the synaptic strengths JE and JI . For instance, lifting θ, or lowering Vr, is equivalent to reducing JE and JI together by the same ratio. Thus, heterogeneity in θ and Vr can be included in µ and σ by means of rescaling. Unlike the above five parameters related to the poten- tial, the remaining two shaping the neuronal response in the temporal scale, τm and τr, cannot be rescaled or cap- tured by µ and σ. Their distributions among neurons have to be accounted for separately. Therefore, in the high input regime when the approximation of a fluctuat- ing input noise is valid, the seven independent neuronal parameters (and their distributions) can be reduced to four: µ, σ, τm and τr. Based on this analysis, we suggest using distributed values of these four parameters together with an identical noise η(t) to account for the effects of all the parameters in a population of LIF neurons re- ceiving identical inputs. This is in contrast to the com- mon practice of using independent noises as shown in Fig. 2(a). We draw the parameters from uniform distri- butions µ ∈ [0.015, 0.105], σ ∈ [0.1, 0.3], τm ∈ [16, 24] ms In Fig. 2(c), both the rate dif- and τr ∈ [1.5, 2.5] ms. ference and the correlation coefficient, as well as their relation, are consistent with the experimental and our simulation results. The respective contributions of the four parameters are investigated. Fig. 3 shows the correlation as functions of the rate difference for µ, σ, τm and τr separately. Each realization is drawn from a uniform distribution of 10 %, 20 % and 50 % around their mean values, which are 0.06, 0.2, 20 ms and 2 ms, respectively. The firing rate of a neuron is largely shaped by µ, whereas the distributed values of the variance give rise to different degrees of imprecise spiking. The wider the distribution of τm and τr, the larger is the rate difference and the lower the correlation. When τr ≪ 1/ν, the effect of τr is small. 1.0 0.5 a r r o C e k p S i 0.0 0 c 1.0 0.5 r r o C e k p S i 0.0 0 20 m 20 Rate difference (Hz) 3  20 r 50% 20% 10% 40 50% 20% 10% 20 Rate difference (Hz) 40 b 1.0 0.5 50% 20% 10% 0.0 0 40 50% 20% 10% d 1.0 0.5 0.0 0 40 FIG. 3. (Color online). Correlation coefficient as a function of rate difference when neurons receive identical noise with uni- formly distributed (a) µ, (b) σ, (c) τm and (d) τr separately. a 80 60 40 ) z H ( e t a R 20 =0 =0.1 =0.3    b 80 60 40 20 =0 =0.05 =0.1 c 80 60 40 20 0 0 0.05  0.1 0 0 0.1 0.2  0.3 0 10 20 m 30 FIG. 4. (Color online). Output rate as a function of (a) µ, (b) σ and (c) τm from theory (black) and simulation (colored). V. RATE DIFFERENCE The firing rate of a LIF receiving a Gaussian dis- tributed noise is known analytically [13, 14]: = τr + τm√πZ θ−µτm σ√τm Vr −µτm σ√τm θ 1 ν 1 ν dueu2 (cid:0)1 + erf(u)(cid:1), when σ > 0; = τr − τmln(1 − µτm ), when σ = 0. (6) Only six parameters µ, σ, τm, τr, θ and Vr influence the firing rates, of which only the first four are independent. Changing any of them can result in a rate difference as shown in Fig. 4, and this explains the wide distribution of firing rates in a heterogeneous neuronal population. VI. IMPRECISE SPIKING The raster plot in Fig. 2(c) shows population synchrony with spike-time jitter. Low average correlation coeffi- cients in a population of neurons do not necessarily im- ply an asynchronous state of the population. Removing spikes from one of the two identical spike trains can also reduce the correlation coefficient significantly. When we compute the correlation coefficient in our data with a a 1.0 Spike corr b 0.10 rxy(  ) c 1.0  VII. DISCUSSION 4 0.5 0.0 0 0.05 0.00 20 40 60 Bin (ms) 80 100 10 5  0  5 10 0.5 0.0 0 20 60 Rate difference (Hz) 40 80 FIG. 5. (Color online). (a) Correlation coefficient as a func- tion of window size (colored lines indicate five examples; the black thick line indicates the average over all pairs). (b) Cross-correlation. Positive correlation at negative τ indicates that the higher-rate neuron is leading. (c) κ, which takes jittered spikes into account, as a function of rate difference. larger bin size, larger values for the the correlation co- efficient are obtained, as shown in Fig. 5(a). This is be- cause some spikes become "coincident" only for larger bin sizes. The output spike trains behave like jittered spikes as discussed in earlier theoretical studies [15]. Whether the decorrelation due to neuronal heterogeneity is signifi- cant depends critically on the bin size, or the integration window of the neurons receiving such inputs. In view of a significant number of spikes jittering out- side a 1 ms bin, we look into the cross-correlation function rxy(τ ) = cxy(τ ) σxσy = hx(t)y(t + τ )i − hx(t)ihy(t)i σxσy , (7) where x(t) and y(t) denote two output spike trains in dis- crete time, consisting of 0 and 1 with bin size of 0.1 ms. x(t) is assigned to be the spike train with lower spike count. cxy(τ ) is the covariance function and σx and σy denote the standard deviation of the two spike trains, considered as discrete signals. Fig. 5(b) shows that the mean rxy(τ ) over all pairs is positive in a small neighbor- hood of τ , indicating a higher than chance level to ob- serve spikes. Spikes are jittered, instead of asynchronous. In addition, rxy(τ ) is asymmetric, and its area is skewed towards negative τ , indicating that the higher-firing-rate neuron is more likely to lead in terms of spiking [16, 17]. the normalized total cross- We further look at covariance κ [15, 18] We remark that the plots of the correlation coefficient as a function of rate difference from our simulations (such as Figs. 1(b) and 2(c) can fit the experimental results in [3] more satisfactorily by introducing a random delay of up to 1 ms to every incoming input spike. The spike correlation distribution then becomes broader, and the regime of low correlated output at small rate differences can also be reached (data not shown). This effect could be due to some (unknown) biological mechanism not cap- tured in a simple LIF neuron model. We emphasize that common input into heterogeneous neurons is better realized by a shared noise with dis- tributed mean and variance and, more completely, with additionally distributed values of the membrane time constant and the refractory period. This insight is based on both the mathematical analysis presented here and the in vitro experimental findings [3]. As far as firing rates and spike correlations are concerned, the distribu- tions of mean and variance of the input account for most of the experimental observations. Diversity of τm and τr has a smaller effect on the quantities in question, never- theless, including them can account for the full degree of heterogeneity. In the raster plot of a neuronal population with sim- ilar rate differences and spike correlations, synchrony is obvious, although spike times are not precise. Spikes, if present, are jittered in the millisecond range, which can- not be captured by the 1 ms temporal window used for analysis. This is why using larger bins leads to larger values for spike correlation. We emphasize that neu- ronal heterogeneity alone does impose an appreciable decorrelation effect on the population activity. However, whether decorrelation is functionally significant depends on the readout of the downstream neurons. On top of that, a network of heterogeneous neurons may give rise to richer network dynamics. It remains to be explored whether such intrinsic heterogeneity can facilitate other decorrelation mechanisms to increase the amount of in- formation flow [3, 19, 20]. Given the significant reduc- tion in spike correlation among heterogeneous neurons, research concerning correlation transmission must take neuronal heterogeneities into consideration. κ = −∞ dτ cxy(τ ) R ∞ qR ∞ −∞ dτ cxx(τ )R ∞ −∞ dτ ′cyy(τ ′) , (8) ACKNOWLEDGMENTS which is an overall measure for the fraction of the spikes that are correlated above chance level. Fig. 5(c) shows that κ is quite close to unity, and has a weak depen- dence on the rate difference. This indicates that spikes would not be decorrelated when a larger time window is considered. We thank Volker Pernice, Moritz Deger, and Tom Tet- zlaff for discussions. The present work was supported by the German Federal Ministry of Education and Research (BMBF Grant No. 01GQ0420 to "BCCN Freiburg" and BMBF Grant No. 01GW0730 "Impulse Control") and the EU (INTERREG-V Grant to Neurex: TriNeuron). [1] E. Marder and J.-M. Goaillard, Nat. Rev. Neurosci. 7, [2] Z. F. Mainen and T. J. Sejnowski, Science 268, 1503 563 (2006). (1995). (2008). [3] K. Padmanabhan and N. N. Urban, Nat. Neurosci. 13, [13] N. Brunel and V. Hakim, Neural Comput. 11, 1621 1276 (2010). (1999). [4] M. I. Chelaru and V. Dragoi, PNAS 105, 16344 (2008). [5] J. Mejias and A. Longtin, Phys. Rev. Lett. 108 (2012). [6] M. Tsodyks, and H. Sompolinsky, I. Mitkov, [14] A. Siegert, Phys. Rev. 81, 617 (1951). [15] T. Tetzlaff, S. Rotter, E. Stark, M. Abeles, A. Aertsen, and M. Diesmann, Neural Comput. 20, 2133 (2008). 5 [7] X. Phys. Rev. Lett. 71, 1280 (1993). and J. Neurosci. 16(20), 6402 (1996). Wang J. [8] R. Brette, PLoS Comput. Biol. 8, e1002561 (2012). [9] L. Ricciardi and M. L. Sacerdote, G. Buzs´aki, [16] S. Ostojic, N. Brunel, J. Neurosci. 29, 10234 (2009). and V. Hakim, [17] T. Tchumatchenko, A. Malyshev, T. Geisel, M. Volgu- shev, and F. Wolf, Phys. Rev. Lett. 104, 058102 (2010). and W. T. Newsome, [18] W. Bair, E. Zohary, Biol. Cybern. 35, 1 (1979). J. Neurosci. 21(5), 1676 (2001). [10] A. Kuhn, A. Aertsen, and S. Rotter, J. Neurosci. 24, 2345 (2004). [11] J. de la Rocha, B. Doiron, E. Shea-Brown, K. Josi´c, and A. Reyes, Nature 448, 802 (2007). [19] A. Renart, J. de la Rocha, P. Bartho, L. Hollender, N. Parga, A. Reyes, and K. D. Harris, Science 327, 587 (2010). [20] M. T. Wiechert, B. Judkewitz, H. Riecke, and R. W. [12] J.-n. Teramae and T. Fukai, Phys. Rev. Lett. 101, 248105 Friedrich, Nat. Neurosci. 13, 1003 (2010).
1509.01913
3
1509
2016-02-18T08:20:46
Fractal Fluctuations in Human Walking: Comparison between Auditory and Visually Guided Stepping
[ "q-bio.NC" ]
In human locomotion, sensorimotor synchronization of gait consists of the coordination of stepping with rhythmic auditory cues (auditory cueing, AC). AC changes the long-range correlations among consecutive strides (fractal dynamics) into anti-correlations. Visual cueing (VC) is the alignment of step lengths with marks on the floor. The effects of VC on the fluctuation structure of walking have not been investigated. Therefore, the objective was to compare the effects of AC and VC on the fluctuation pattern of basic spatiotemporal gait parameters. Thirty-six healthy individuals walked 3 x 500 strides on an instrumented treadmill with augmented reality capabilities. The conditions were no cueing (NC), AC, and VC. AC included an isochronous metronome. In VC, projected stepping stones were synchronized with the treadmill speed. Detrended fluctuation analysis assessed the correlation structure. The coefficient of variation (CV) was also assessed. The results showed that AC and VC similarly induced a strong anti-correlated pattern in the gait parameters. The CVs were similar between the NC and AC conditions but substantially higher in the VC condition. AC and VC probably mobilize similar motor control pathways and can be used alternatively in gait rehabilitation. However, the increased gait variability induced by VC should be considered.
q-bio.NC
q-bio
Fractal Fluctuations in Human Walking: Comparison between Auditory and Visually Guided Stepping PHILIPPE TERRIER 1,2 1IRR, Institute for Research in Rehabilitation, Sion, Switzerland; 2Clinique romande de réadaptation SUVACare, Sion, Switzerland Address correspondence to Philippe Terrier, Clinique romande de réadaptation SUVACare, Av. Gd-Champsec 90, 1951 Sion Switzerland. Tel.: +41-27-603-20-77. Electronic mail: Philippe.Terrier[at]crr-suva.ch. Notice: this is the author's version of a work that was accepted for publication in the Annals of Biomedical Engineering. Changes resulting from the publishing process, such as final peer review, editing, corrections, structural formatting, and other quality control mechanisms are not reflected in this document. The final publication is available at Springer via http://dx.doi.org/10.1007/s10439- 016-1573-y . Fractal fluctuations of guided stepping 2 Abstract-In human locomotion, sensorimotor synchronization of gait consists of the coordination of stepping with rhythmic auditory cues (auditory cueing, AC). AC changes the long-range correlations among consecutive strides (fractal dynamics) into anti-correlations. Visual cueing (VC) is the alignment of step lengths with marks on the floor. The effects of VC on the fluctuation structure of walking have not been investigated. Therefore, the objective was to compare the effects of AC and VC on the fluctuation pattern of basic spatiotemporal gait parameters. Thirty-six healthy individuals walked 3 × 500 strides on an instrumented treadmill with augmented reality capabilities. The conditions were no cueing (NC), AC, and VC. AC included an isochronous metronome. In VC, projected stepping stones were synchronized with the treadmill speed. Detrended fluctuation analysis assessed the correlation structure. The coefficient of variation (CV) was also assessed. The results showed that AC and VC similarly induced a strong anti-correlated pattern in the gait parameters. The CVs were similar between the NC and AC conditions but substantially higher in the VC condition. AC and VC probably mobilize similar motor control pathways and can be used alternatively in gait rehabilitation. However, the increased gait variability induced by VC should be considered. Key terms-Human locomotion, Motor control, Sensorimotor synchronization, Gait variability, Auditory cueing, Visual cueing, Long-range correlations. Abbreviations: AC: Auditory cueing NC: No cueing VC: Visual cueing DFA: Detrended fluctuation analysis CV: Coefficient of variation Author's accepted version Rev2, Feb. 2016 Fractal fluctuations of guided stepping 3 INTRODUCTION In the 1970s, Benoit Mandelbrot laid the foundations of a new method for understanding the geometry of nature. He coined the term "fractal" to describe geometric objects that look identical, regardless of the scale at which they are observed (self-similarity).20 He also developed an analogous concept about particular time series that present self-similarities.21 In this case, a constant statistical distribution exists across time scales. In other words, the statistical features of the parts of the series are comparable, even if the time interval during which the observations are made changes. The corollary is that these fractal time series exhibit serial correlations (or autocorrelations) between consecutive samples that slowly decrease under a power law (long-range correlations). Biological systems are inherently complex. They are constituted of multiple sub-parts-ranging from the molecular to the population level-that interact nonlinearly on large spatial and temporal scales. Consequently, signals measured from living organisms are most often fractal. Although the classical approach considers the fluctuations in a physiological signal as random noise around the true average, the fractal analysis postulates that the structure of fluctuations enlightens us about the underlying processes that produced the signal.10 Fractal analysis has been used on a wide variety of signals, such as heartbeat time series29 or electroencephalograms.1 In human walking, the muscles of the lower limbs cyclically propel the body forward over a certain distance (step length) during a certain time (step time), thus maintaining ambulatory speed (length/time). Gait control expends low energy by delivering an optimal combination of step length and step time.48 Furthermore, an active control of lateral foot placement is needed to provide a base of support that minimizes fall risks.3 In short, the control of human walking is a highly complex process implying intricate interactions of feed-forward and feed-back mechanisms,11, 35 which is a condition that is conducive for the emergence of fractal patterns in gait fluctuations. Actually, in 1995, by analyzing the walk of healthy individuals, Hausdorff and colleagues observed that the time series of stride time (i.e., the time between two consecutive heel strikes of the same limb) exhibited fractal fluctuations16: Deviations above and below the average tend to persist over several decades of later strides (long memory process). In order to highlight the fractal patterns, Hausdorff et al. used detrended fluctuation analysis (DFA), which was designed to assess the scaling properties of time series with nonstationarities.28 In subsequent studies using DFA, they observed that, in patients with neurological gait disorders, the fractal pattern tends to be replaced by a random pattern (i.e., no correlations among successive strides).12, 14 Thus, DFA has often been adopted to study locomotor function in patients with neurological disorders.13 Author's accepted version Rev2, Feb. 2016 Fractal fluctuations of guided stepping 4 During the decade after these seminal works, further studies using DFA extended the knowledge about fractal fluctuations in human locomotion. In 2003, West and Scaffetta46 proposed a nonlinear stochastic dynamical model of walking that accounted for the presence of long-range correlations in stride time series. In 2005, Terrier et al.42 observed that the time series of stride time, as well as the time series of stride length and stride speed were fractal. The presence of fractal patterns in several spatiotemporal gait parameters was later confirmed by Jordan et al. in a treadmill experiment.18 In the meantime, theoretical considerations highlighted some issues with the DFA method,22 questioning the presence of true long-range correlations in gait time series. However, in 2009, Delignières and Torre5 tested the effective presence of long-range correlations in human walking with an alternative methodology43: by comparing different synthetic signals with actual stride time series, they concluded that artificial time series that incorporated long-range correlation by design (ARFIMA models) best fit human time series. An interesting discovery has been that sensorimotor synchronization substantially alters the fractal structure of human walking. Sensorimotor synchronization is the coordination of movements with external rhythms or cues. This ability is responsible for unique human behavior, such as dancing or performing music.32 In locomotion, sensorimotor synchronization implies guided stepping with external cues. The simplest expression of this synchronization behavior consists of the voluntary adjustment of heel strikes to the beat of an isochronous metronome, hereafter simply referred to as auditory cueing (AC). AC is important in gait rehabilitation. For example, in stroke patients, a recent meta-analysis of seven randomized controlled trials has concluded that the gait training with AC improves walking speed, stride length and cadence.25 AC is also a key tool for improving gait among patients with Parkinson's disease.45 AC has a strong effect on the fractal fluctuations of stride time but no effect on stride length and stride speed.36, 42 The long-range correlated pattern is then replaced by anti-correlations. A value above the mean is more likely to be followed by a value below the mean (anti-persistence): the voluntary control of step duration induces a continual shift around the target (over-correction).8, 41 Another type of externally-driven synchronization of human movement is treadmill walking. Indeed, a motorized treadmill imposes a constant speed upon the walker. Whereas treadmill walking has only a marginal effect on the fractal fluctuation of stride time compared to overground walking,4, 39 treadmill walking changes the serial correlations of stride speed into anti-correlations, as AC does for stride time.8, 34, 41 Moreover, when AC is combined with treadmill walking, all the gait parameters (stride time, stride length, and stride speed) exhibit anti-correlated patterns.34, 41 The constant speed Author's accepted version Rev2, Feb. 2016 Fractal fluctuations of guided stepping 5 of the treadmill (speed cueing) requires coordinated adjustments between the stride time and the stride length to maintain an appropriate speed and thus avoid falling off the treadmill. As a result, when cadence and speed are voluntarily adjusted to external cues (dual cueing), stride length must be coordinated accordingly (the "loss of redundancy" paradigm).8, 34, 41 In addition to the temporal adjustments of steps to auditory cues, another possible type of sensorimotor synchronization consists of adjusting step length to visual cues (hereafter referred to as visual cueing, VC). In this case, a walker anticipates the position of his or her next step to coincide with a mark on the floor. Like AC, VC has applications in gait rehabilitation. In Parkinson's disease, gait training with VC might have a long-term beneficial effect on walking capabilities.37 Recent technical advances have led to the development of treadmills equipped with projection devices that draw visual targets on the treadmill belt (augmented reality). The use of projected visual targets on a treadmill resembles the more conventional, real, marks on the ground, but greatly facilitates the application of VC for research and rehabilitation. Promising results have been obtained with such treadmills, in particular for stroke rehabilitation.17 In patients in the chronic stage of stroke, adaptability training using visually guided stepping improved obstacle-avoidance performance.44 Although the field is still in its infancy, by offering complementary solutions to AC, the VC method has substantial potential for growth. In summary, numerous studies have analyzed the effects of AC and treadmills on the variability8, 15, 38, 39 and the fractal pattern5, 34, 42 in human locomotion. This has led to interesting hypotheses about the neurological basis of gait control, such as the existence of a specific neural structure that generates fractal noise (the super central pattern generator hypothesis),5, 46 or the implication of the minimum intervention principle.8, 41 However, whether or not VC walking supports these hypotheses has yet to be investigated. The first objective of the present study was to determine, in healthy individuals walking on a treadmill, the effects of VC on stride-to-stride fluctuations and to compare the results with the effects of AC. The hypothesis was that VC and AC similarly alter the fractal fluctuations normally present in the time series of stride time and stride length (loss of redundancy) and that the long- range correlations are assumed to be replaced by anti-correlations. The second objective was to measure the fluctuation magnitude (the coefficient of variation, CV) of the spatiotemporal gait parameters under different cueing conditions. Author's accepted version Rev2, Feb. 2016 Fractal fluctuations of guided stepping 6 MATERIALS AND METHODS Subjects Thirty-six healthy volunteers (14 men, 22 women) with no orthopedic or neurological problems participated in the study. The mean and standard deviation (SD) of their characteristics were as follows: age 33 years (10), body height 1.72 m (0.08), and body mass 66 kg (13). The subjects had had no previous experience with walking following visual cues. All subjects signed an informed consent form according to the guidelines of the local ethic committee (Commission Cantonale Valaisanne d'Ethique Médicale, [CCVEM]), that had approved the protocol. Material and Experimental Procedure The instrumented treadmill was a C-mill model (ForceLink BV, Culemborg, The Netherlands),17 which is equipped with an embedded vertical force platform of 70 × 300 cm. The platform recorded the vertical force and the position of the center of pressure at a sampling rate of 500 Hz. A projection system displayed visual objects on the walking area from the right side of the treadmill. Ad hoc control software (CueFors®) was used to compute the preliminary values of basic gait parameters (stride length and duration) and to control the projection of the visual cue drawings, the stepping stones."17 During all procedures, an elastic band was placed in front of the participant (1.40 m behind the beginning of the belt), hanging perpendicular to the handrails, at hip level. The participant was instructed to stay 10 cm behind the band, ensuring consistent placement on the walking area (approximately in the middle). The reasons were the increased safety and the standardization of the number of incoming stepping stones seen by the participant. Firstly, the preferred walking speed (PWS) of each participant was assessed using a standardized procedure,9 which consisted of (1) a progressive increase in the treadmill speed from a low speed (2 km h–1) until the participant reported a comfortable pace and then (2), similarly, a progressive decrease of the treadmill speed from a high speed (6 km h–1) to a comfortable pace. The PWS was defined as the average of four tests: two with increasing speeds and two with decreasing speeds. Then, the participant performed a 2 min walking test at PWS. The last 30 steps were analyzed to measure the average preferred stride length and stride time (i.e., the duration of one gait cycle). Then, the participant walked continuously for about 30 min at his or her PWS. Three conditions were presented in a random order. Before each condition, the experimenter described the task to the Author's accepted version Rev2, Feb. 2016 Fractal fluctuations of guided stepping 7 subject while he or she continued to walk. The conditions were as follows: (1) no cueing, i.e., normal walking (NC); (2) AC, i.e., walking while synchronizing the heel strike to the beep of an electronic metronome set to the preferred previously measured cadence; (3) VC, walking while adjusting steps to stepping stones, which were 20 × 30 cm moving rectangles projected on the walking area that went back at the same speed as the treadmill belt. The longitudinal distance along the belt between the successive stepping stones was set to correspond to half of the preferred stride length. The instruction was to aim accurately for each stepping stone with the foot. Given his or her position in the middle of the treadmill, the participant could see in advance two stepping stones. For a better understanding of the VC method, see the short movie provided in the supplemental material (S1). Thirty seconds of familiarization with the cueing task was given to the volunteer before the recording began. One thousand steps (500 gait cycles) were then recorded, which correspond to 10 min of walking at a step rate of 100 steps min–1. Data Analysis One hundred and eight files (36 subjects × 3 conditions), each containing the position of the center of pressure recorded at 500 Hz, were exported for the subsequent analyses that were performed using MATLAB (MathWorks, Natick, MA). The stride time and length of each gait cycle (500 per file) were computed using a custom algorithm that implemented a validated method.33 The reasoning behind the method is to detect heel strikes in the longitudinal signal and then to compute the distance and time between subsequent heel strikes of the same foot taking into account the treadmill speed. The average speed of each gait cycle was defined as speed = stride length/stride time. As a result, three time series (NC, AC, VC) of 500 gait cycles were obtained for stride time, stride length, and stride speed (total 3 × 3 × 36 = 324). A typical result for one participant is shown in Fig. 1. To characterize the dispersion of the values around the mean in the time series (fluctuation magnitude), the CVs were computed, which were defined as the SD normalized by mean, and were expressed in percentages. A fractal time series with long-range correlations exhibits an autocorrelation function C(s) that declines following a power law (cid:2)(cid:3)(cid:4)(cid:5)∝ (cid:4)(cid:7)(cid:8),0<(cid:12)<1. The DFA is a method designed to assess the scaling exponent α (α = 1 – γ / 2) in a time series with nonstationarities.29 Although DFA may not always be appropriate to evidence the existence of long-range correlations in time series5, 22, 43, here, I took advantage of its proved capacity to efficiently distinguish between statistical persistence Author's accepted version Rev2, Feb. 2016 Fractal fluctuations of guided stepping 8 (short- or long-range correlations) and anti-persistence (anti-correlations).7, 8, 41 First, the time series of length N was integrated. Then, it was divided into non-overlapping boxes of equal length n. In each box, a linear fit, using the least squares method, was performed. The average fluctuation F(n) for that box length (cid:14)(cid:3)(cid:15)(cid:5)=(cid:17)1(cid:18)(cid:19)(cid:20)(cid:21)(cid:3)(cid:22)(cid:5)−(cid:21)(cid:24)(cid:3)(cid:22)(cid:5)(cid:25)(cid:26) (cid:27) (cid:28)(cid:29)(cid:30) was: where yn(k) was the y-coordinate of the kth point of the straight line resulting from the linear fit, and y(k) was the corresponding point in the original time series. The procedure was repeated for increasing box sizes. Box sizes (n) ranging from 12 to 125 (i.e., N / 4) were used, with exponential spacing to avoid a bias toward larger box sizes.23 A statistical self-affinity at different scales implies that F(n) » nα . Therefore, because logF(n) » α·log(n), a linear fit is realized in a log-log plot between n and F(n) to compute the scaling exponent α. If α lies between 0.5 and 1, a long-range correlation is likely. A random process (white noise) induces α value of 0.5. In the case of anti-correlation, a small α is expected (α < 0.5).40 Statistics Six dependent variables were included in the statistical analysis, namely the CV and the scaling exponent α for each gait parameter: stride time, stride length and stride speed. The independent variable was the cueing condition (NC, AC, and VC). Boxplots were used to describe the distribution of the individual results (Figs. 2 and 3). Means and SDs are shown in Tables 1 and 2. As inferential statistics, one-way, repeated-measures analyses of variance (ANOVA) were used. The explained variance was assessed with partial η2. Post hoc analyses with Tukey's honestly significant difference (HSD) tests were used to highlight specific differences between the cueing conditions; only the p- values of the significant differences have been included in Tables 1 and 2, with associated relative differences. RESULTS Fluctuation Magnitude / Gait Variability The distribution of the variability results (CV) is presented in Fig. 2. The presence of some outliers can be observed. A substantially higher CV of stride time, length, and speed is observed for the VC Author's accepted version Rev2, Feb. 2016 Fractal fluctuations of guided stepping 9 condition. Inferential statistics (Table 1) confirm that cueing had a significant effect on the CV of the gait parameters. The ANOVA results show that cueing conditions explain a substantial part of the variance (partial η 2: 0.56–0.72). In other words, VC induced a substantial increase in the gait parameters' CV compared to NC, whereas for AC such an increase in fluctuation magnitude was not observed. Fractal Analysis The boxplots of the fractal analysis results (DFA) show that stride time and stride length exhibited long-range correlations under NC conditions (α > 0.5), while stride speed was constantly anti- correlated (α < 0.5). An obvious anti-correlated pattern was observed for all gait parameters in the AC and VC conditions. The ANOVA results showed significant differences among the cueing conditions for all gait parameters, with the exception of stride speed. Concerning stride length and stride time, a very large part of the variance is explained by the model (partial η 2 > 0.79), due to the change in the scaling exponent from correlation to anti-correlation. Post hoc analyses confirmed large changes in the correlation structure for stride length and stride time when AC and VC are compared to NC. For these parameters, no difference existed between the AC and VC conditions (similar anti-correlations). DISCUSSION Based on the analysis of a large number of gait cycles (54,000), the objective of the present study was to characterize the effects of external cues on the stride-to-stride fluctuations in gait parameters, in structural (fractal pattern) and magnitude (CV) terms. As hypothesized, concerning fractal fluctuations, spatiotemporal parameters responded similarly to AC and VC, namely, the emergence of strong anti-correlations was observed. However, AC and VC did not affect the fluctuation magnitude in the same way: VC induced a substantial increase in the gait parameters' CVs compared to the control condition, whereas no differences were observed for AC. When individuals walk in a constant environment, basic spatiotemporal parameters fluctuate from stride-to-stride in a narrow window, very likely framed by energetic, biomechanical,26, 48 and stability27 constraints. In overground walking, typical CV values between 2.5% and 3% have been reported in healthy young subjects.39, 42 Similar results have been observed in treadmill walking,39, 41 which were confirmed in the present study (CV 1.8% to 2.8%, Table 1). The effect of AC on the Author's accepted version Rev2, Feb. 2016 Fractal fluctuations of guided stepping 10 fluctuation magnitude has been assessed in experimental and clinical studies. In elderly subjects, in two studies, which analyzed overground walking with or without AC given at a preferred cadence, a small increase in gait variability15 or no effect47 was found. In a previous study that combined treadmill and AC (N = 20 young adults, 5 min walking), when the AC and NC conditions were compared, a small decrease in the CVs for the stride time (–19%), stride length (–17%), and stride speed (–7%) has been observed.41 Compared to that study, the present study included more participants (N = 36) and longer trials (10 min). The (not significant) changes induced by AC were stride time, –13%; stride length, –7%; and stride speed, 0% (Table 1). In summary, AC has little impact on stride-to-stride fluctuation magnitude. As illustrated in Fig.1, the erratic wandering of the parameter around the mean, which is typical of fractal fluctuation (first column), is replaced by high frequency noise around the imposed frequency, which is typical of anti-correlated patterns (second column). Thus, these concomitant changes in the fluctuation pattern triggered by AC, which decreased low frequency wanderings and increased high frequency oscillations, nearly compensate for each other to give a comparable fluctuation magnitude. On the other hand, VC had a profound effect on the fluctuation magnitude. The stride time CV increased by 51%, stride length CV by 73%, and stride speed CV by 63% (Table 1). Although the participants were able to target the stepping stones, their stride length oscillated in an extended range (CV: 5%, 7 cm for a 1.3 m stride length). Due to the interdependence among gait parameters, the stride time and the stride speed were also affected. This suggests that VC was a challenging task. A cause of the difficulty was perhaps that only two incoming steps were seen in advance by the subject, which could be too short a warning to anticipate precise foot placement. Peper et al.30, using the same treadmill as in the present study, showed that individuals spontaneously chose a lower walking speed in VC condition: this corroborates the hypothesis that a sufficient reaction time to the incoming targets should be allowed to help the subject to be comfortable with VC. However, further studies are needed to analyze gait variability at different speeds under VC condition. As far as I know, this strong effect on fluctuation magnitude when VC is combined with treadmill walking has not been described in the literature and has to be confirmed independently. Interestingly, high variability has been observed in the pathological gait of patients with Parkinson's disease (CV of stride length: 5.32%).19 Patients with Parkinson's disease rely more on vision while walking.2 The fact that healthy individuals, who guide their steps based on vision, also exhibited high variability may have important implications in fundamental research and clinical application, which deserve further investigations. Walking on a treadmill requires coordinated control of stride time and stride length to match the treadmill speed. As evidence of this coordinated regulation, several studies have shown that stride Author's accepted version Rev2, Feb. 2016 Fractal fluctuations of guided stepping 11 time and stride length are actually cross-correlated:7, 34, 41 In other words, stride time and stride length vary over time in a similar way (positive correlations).34 A model that explains how motor control manages speed regulation has been proposed by Dingwell and collaborators.6, 8 A key feature of this model is that an infinite combination of stride time and stride length is possible to meet the goal of maintaining a constant speed. In other words, redundancy exists between spatial and temporal control of walking speed: if a deviation occurs in the stride length, the deviation can be compensated for by a correction in the stride time, and vice versa. As a result, deviations can persist over consecutive strides, which may explain long-range correlations and fractal fluctuations (the minimum intervention principle).6 If two goals are simultaneously imposed on a walking individual, for instance, treadmill walking (speed goal) and AC (stride time goal), the redundancy disappears, and tight regulation of stride length is also performed, which leads to the loss of the fractal fluctuations, as demonstrated in previous studies8, 34, 41 and in the present study (Fig. 3). The main goal of the present study was to provide further evidence to support this model. As expected, imposing a spatial goal (stepping to visual cues) modified the fluctuation pattern of stride length as well as stride time (Fig. 3 and Table 2). Actually, AC and VC altered fractal fluctuations in a similar way (Table 2). As observed previously, strong anti-correlations appeared.41, 42 As illustrated in Table 3, the results of the present study provide new information to complement what is already known about voluntary synchronization of gait to external cues. The main gap that remains to be filled is the fluctuation pattern of overground walking under the VC condition. It can be assumed that stride length would then be anti-correlated, while stride time and stride speed would remain correlated. However, conducting such an experiment is technically challenging. Taken together, these findings highlight a stereotyped response of gait control to external cueing, the neurobiological basis of which remains to be elucidated. What is already known is that sensorimotor synchronization requires anticipation to align motor response with external cues. The delivery of sensory signals and central processing delay the motor response. Thus, to get in time, motor command must anticipate future steps based on internal models and past sensory inputs and not react to current stimuli.24 This anticipation leads to a well-known phenomenon in rhythmic tapping experiments: the taps precede the metronome (negative mean asynchrony).31 Likewise, in gait experiments with AC, the heel-strike slightly precedes (50 ms on average) the next occurrence of the metronome sound.23 The anticipation of movement based on visual cues, for example, reaching a moving object, is also a basic task of sensorimotor coordination that implies feedforward mechanisms.24 In the present experiment, the subject saw two stepping stones in advance to Author's accepted version Rev2, Feb. 2016 Fractal fluctuations of guided stepping 12 anticipate the correct step lengths. In short, guided stepping requires voluntary adaptation to external cues and anticipation to produce a timed motor response. It can be assumed that visual, auditory, and somatosensory afferents converge to modulate a central pacemaker that triggers the appropriate rhythmic gait behavior under voluntary control. How exactly this sensory feedback feeds this hypothetical pacemaker, and how an anti-correlated pattern is produced remain to be further investigated. In conclusion, because it can be assumed that AC and VC mobilize the same motor pathways, they can probably be used alternatively in gait rehabilitation. The efficiency of VC to enhance walking abilities in patients with neurological gait disorders needs further studies. However, the high gait variability induced by VC might have detrimental effects, for instance, a lower dynamic balance. This should be taken into account in the development of VC rehabilitation methods. ACKNOWLEDGMENTS The author warmly thanks Emilie Du Fay de Lavallaz for his valuable support in bibliographical research and Vincent Bonvin for his help in data collection. The study was funded by the Swiss accident insurance company SUVA, an independent, non-profit company under public law, and by the clinique romande de réadaptation. The IRR (Institute for Research in Rehabilitation) is funded by the State of Valais and the City of Sion. Study funders had no role in the collection, analysis, and interpretation of data; in the writing of the manuscript; and in the decision to submit the manuscript for publication. Author's accepted version Rev2, Feb. 2016 Fractal fluctuations of guided stepping 13 REFERENCES Azulay J.-P., S. Mesure, B. Amblard, O. Blin, I. Sangla and J. Pouget. Visual control of Dingwell J. B. and J. P. Cusumano. Identifying Stride-To-Stride Control Strategies in Human Dingwell J. B. and J. P. Cusumano. Re-interpreting detrended fluctuation analyses of stride- Dingwell J. B., J. John and J. P. Cusumano. Do humans optimally exploit redundancy to control Dingwell J. B. and L. C. Marin. Kinematic variability and local dynamic stability of upper body Eke A., P. Herman, L. Kocsis and L. Kozak. Fractal characterization of complexity in temporal Chang M. D., S. Shaikh and T. Chau. Effect of treadmill walking on the stride interval dynamics Accardo A., M. Affinito, M. Carrozzi and F. Bouquet. Use of the fractal dimension for the Bauby C. E. and A. D. Kuo. Active control of lateral balance in human walking. J. Biomech. Delignieres D. and K. Torre. Fractal dynamics of human gait: a reassessment of the 1996 data 1. analysis of electroencephalographic time series. Biol. Cybern. 77:339-350, 1997. 2. locomotion in Parkinson's disease. Brain 122:111-120, 1999. 3. 33:1433-1440, 2000. 4. of human gait. Gait Posture 30:431-435, 2009. 5. of Hausdorff et al. J. Appl. Physiol. 106:1272-1279, 2009. 6. Treadmill Walking. PloS ONE 10:e0124879, 2015. 7. to-stride variability in human walking. Gait Posture 32:348-353, 2010. 8. step variability in walking? PLoS Comput. Biol. 6:e1000856, 2010. 9. motions when walking at different speeds. J. Biomech. 39:444-452, 2006. 10. physiological signals. Physiol. Meas. 23:R1, 2002. 11. a systematic review. Neurosci. Biobehav. R. 57:310-327, 2015. 12. stride length, gait variability, and fractal-like scaling. Chaos 19:026113, 2009. 13. fluctuations of human walking. Hum. Mov. Sci. 26:555-589, 2007. 14. Hausdorff J. M., A. Lertratanakul, M. E. Cudkowicz, A. L. Peterson, D. Kaliton and A. L. Goldberger. Dynamic markers of altered gait rhythm in amyotrophic lateral sclerosis. J. Appl. Physiol. 88:2045-2053, 2000. 15. Hausdorff J. M., J. Lowenthal, T. Herman, L. Gruendlinger, C. Peretz and N. Giladi. Rhythmic auditory stimulation modulates gait variability in Parkinson's disease. Eur. J. Neurosci. 26:2369-2375, 2007. 16. Hausdorff J. M., C. K. Peng, Z. Ladin, J. Y. Wei and A. L. Goldberger. Is walking a random walk? Evidence for long-range correlations in stride interval of human gait. J. Appl. Physiol. 78:349-358, 1995. Heeren A., M. W. van Ooijen, A. C. Geurts, B. L. Day, T. W. Janssen, P. J. Beek, M. Roerdink 17. and V. Weerdesteyn. Step by step: a proof of concept study of C-Mill gait adaptability training in the chronic phase after stroke. J. Rehabil. Med. 45:616-622, 2013. 18. Gait Posture 26:128-134, 2007. Hamacher D., F. Herold, P. Wiegel, D. Hamacher and L. Schega. Brain activity during walking: Hausdorff J. M. Gait dynamics in Parkinson's disease: common and distinct behavior among Jordan K., J. H. Challis and K. M. Newell. Walking speed influences on gait cycle variability. Hausdorff J. M. Gait dynamics, fractals and falls: finding meaning in the stride-to-stride Author's accepted version Rev2, Feb. 2016 Fractal fluctuations of guided stepping 14 Lewis G. N., W. D. Byblow and S. E. Walt. Stride length regulation in Parkinson's disease: the O'Connor S. M., H. Z. Xu and A. D. Kuo. Energetic cost of walking with increased step Peng C. K., S. Havlin, H. E. Stanley and A. L. Goldberger. Quantification of scaling exponents Peper C. L. E., M. J. de Dreu and M. Roerdink. Attuning one's steps to visual targets reduces Repp B. H. Sensorimotor synchronization: a review of the tapping literature. Psychon. Bull. Maraun D., H. W. Rust and J. Timmer. Tempting long-memory - on the interpretation of DFA Marmelat V., K. Torre, P. J. Beek and A. Daffertshofer. Persistent fluctuations in stride Owings T. M. and M. D. Grabiner. Variability of step kinematics in young and older adults. Peng C.-K., S. V. Buldyrev, S. Havlin, M. Simons, H. E. Stanley and A. L. Goldberger. Mosaic Miall R. C. and D. M. Wolpert. Forward models for physiological motor control. Neural Mandelbrot B. B. Les objets fractals: forme, hasard et dimension. Flammarion, 1975. Mandelbrot B. B. and J. W. Van Ness. Fractional Brownian motions, fractional noises and 19. use of extrinsic, visual cues. Brain 123 ( Pt 10):2077-2090, 2000. 20. 21. applications. SIAM review 10:422-437, 1968. 22. results. Nonlinear Proc. Geoph. 11:495-503, 2004. 23. intervals under fractal auditory stimulation. PLoS ONE 9:e91949, 2014. 24. networks 9:1265-1279, 1996. 25. Nascimento L. R., C. Q. de Oliveira, L. Ada, S. M. Michaelsen and L. F. Teixeira-Salmela. Walking training with cueing of cadence improves walking speed and stride length after stroke more than walking training alone: a systematic review. J. Physiother. 61:10-15, 2015. 26. variability. Gait Posture 36:102-107, 2012. 27. Gait Posture 20:26-29, 2004. 28. organization of DNA nucleotides. Phys. Rev. E 49:1685, 1994. 29. and crossover phenomena in nonstationary heartbeat time series. Chaos 5:82-87, 1995. 30. comfortable walking speed in both young and older adults. Gait Posture 41:830-834, 2015. 31. Rev. 12:969-992, 2005. 32. 2012). Psychon. Bull. Rev. 20:403-452, 2013. 33. detection using a large force platform embedded in a treadmill. J. Biomech. 41:2628-2632, 2008. 34. Persistent Fractal Metronome without Falling Off the Treadmill? PLoS ONE 10:e0134148, 2015. 35. Physiol. Rev. 86:89-154, 2006. 36. Temporal Dynamics of Human Gait. PLoS ONE 7:e43104, 2012. 37. training using visual cues in an individual with Parkinson disease. Phys. Ther. 86:186-194, 2006. 38. Cueing. PLoS ONE 7:e47171, 2012. 39. treadmill walking. J. Neuroeng. Rehabil. 8:12, 2011. 40. auditory cueing on local dynamic stability. Front. Physiol. 4:230, 2013. 41. treadmill walking: influence of rythmic auditory cueing. Hum.Mov. Sci. 31:1585-1597, 2012. 42. Terrier P., V. Turner and Y. Schutz. GPS analysis of human locomotion: further evidence for long-range correlations in stride-to-stride fluctuations of gait parameters. Hum. Mov. Sci. 24:97-115, 2005. Sidaway B., J. Anderson, G. Danielson, L. Martin and G. Smith. Effects of long-term gait Terrier P. Step-to-Step Variability in Treadmill Walking: Influence of Rhythmic Auditory Repp B. H. and Y.-H. Su. Sensorimotor synchronization: A review of recent research (2006- Roerdink M., B. H. Coolen, B. H. Clairbois, C. J. Lamoth and P. J. Beek. Online gait event Rossignol S., R. Dubuc and J. P. Gossard. Dynamic sensorimotor interactions in locomotion. Sejdic E., Y. Fu, A. Pak, J. A. Fairley and T. Chau. The Effects of Rhythmic Sensory Cues on the Terrier P. and O. Dériaz. Kinematic variability, fractal dynamics and local dynamic stability of Terrier P. and O. Dériaz. Nonlinear dynamics of human locomotion: effects of rhythmic Terrier P. and O. Dériaz. Persistent and anti-persistent pattern in stride-to-stride variability of Roerdink M., A. Daffertshofer, V. Marmelat and P. J. Beek. How to Sync to the Beat of a Author's accepted version Rev2, Feb. 2016 Fractal fluctuations of guided stepping 15 Torre K., D. Delignieres and L. Lemoine. Detection of long-range dependence and estimation van Wegen E. E., M. A. Hirsch, M. Huiskamp and G. Kwakkel. Harnessing Cueing Training for 43. of fractal exponents through ARFIMA modelling. Brit. J. Math. Stat. Psy. 60:85-106, 2007. 44. van Ooijen M. W., A. Heeren, K. Smulders, A. C. Geurts, T. W. Janssen, P. J. Beek, V. Weerdesteyn and M. Roerdink. Improved gait adjustments after gait adaptability training are associated with reduced attentional demands in persons with stroke. Exp. Brain Res. 233:1007-1018, 2015. 45. Neuroplasticity in Parkinson Disease. Top. Geriatr. Rehabil. 30:46-57, 2014. 46. West B. J. and N. Scafetta. Nonlinear dynamical model of human gait. Phys. Rev. E 67:051917, 2003. 47. Wittwer J. E., K. E. Webster and K. Hill. Music and metronome cues produce different effects on gait spatiotemporal measures but not gait variability in healthy older adults. Gait Posture 37:219- 222, 2013. 48. walking. Eur. J. Appl. Physiol. Occup. Physiol. 33:293-306, 1974. Zarrugh M. Y., F. N. Todd and H. J. Ralston. Optimization of energy expenditure during level Author's accepted version Rev2, Feb. 2016 Fractal fluctuations of guided stepping 16 Figures FIGURE 1. Typical results. One participant walked approximately 3X10min on an instrumented treadmill under 3 conditions: no cueing (NC), auditory cueing (AC) and visual cueing (VC). In each condition, 500 strides were recorded (x-axes, #stride). Basic spatiotemporal gait parameters (y-axes), i.e., stride time (ST), stride length (SL), and stride speed (SS), were assessed. Author's accepted version Rev2, Feb. 2016 Fractal fluctuations of guided stepping 17 FIGURE 2. Boxplots of the fluctuation magnitude results. Thirty-six subjects walked on a treadmill under 3 conditions: NC = no cueing (normal walking); AC = auditory cueing (metronome walking); VC = visual cueing (visually guided stepping). In each condition for each subject, 500 gait cycles were recorded, from which the relative fluctuation magnitude (variability) of the spatiotemporal gait parameters was assessed (i.e. CV = SD / mean * 100). Boxplots show the quartiles, the medians and the ranges of the individual results. Outliers are indicated with + signs. Author's accepted version Rev2, Feb. 2016 Fractal fluctuations of guided stepping 18 FIGURE 3. Boxplots of the fractal fluctuation results. Thirty-six subjects walked on a treadmill under 3 conditions: NC = no cueing (normal walking); AC = auditory cueing (metronome walking); VC = visual cueing (visually guided stepping). In each condition for each subject, 500 gait cycles were recorded, from which the serial correlations of the spatiotemporal gait parameters were assessed (scaling exponent α, DFA). Boxplots show the quartiles, the medians and the ranges of the individual results. Outliers are indicated with + signs. Author's accepted version Rev2, Feb. 2016 Fractal fluctuations of guided stepping 19 Tables TABLE 1. Fluctuation magnitude of the gait parameters Descriptive statistics: Mean (SD) ANOVA Post hoc comparisons N=36 NC AC VC F VC vs NC VC vs AC p η2 CI p D % p D % Stride time CV (%) 1.6 (0.6) 1.4 (0.5) 2.3 (0.9) 45.1 <0.001 0.56 0.45 0.70 <0.001 51 <0.001 66 Stride length CV (%) 2.8 (0.8) 2.6 (0.7) 4.8 (1.7) 88.5 <0.001 0.72 0.62 0.83 <0.001 73 <0.001 82 Stride speed CV (%) 1.8 (0.5) 1.8 (0.5) 2.8 (0.9) 83.6 <0.001 0.70 0.60 0.83 <0.001 63 <0.001 62 Thirty-six subjects walked on a treadmill under 3 conditions: NC = no cueing (normal walking); AC = auditory cueing (metronome walking); and VC = visual cueing (visually guided stepping). In each condition for each subject, 500 gait cycles were recorded, from which the relative fluctuation magnitude (variability) of the spatiotemporal gait parameters was assessed (i.e. CV = SD/mean * 100). Descriptive and inferential statistics and significant post hoc comparisons are shown. % is the relative difference: (VC-NC)/NC *100. Significant results are shown in bold. Author's accepted version Rev2, Feb. 2016 Fractal fluctuations of guided stepping 20 TABLE 2. Long-range correlations (fractal fluctuations) of the gait parameters Descriptive statistics: Mean (SD) ANOVA Post hoc comparisons AC vs NC VC vs NC N=36 NC AC VC F p η2 η2 CI p D % p D % Stride time α 0.90 (0.13) 0.32 (0.17) 0.34 (0.14) 209 <0.001 0.86 0.80 0.91 <0.001 -64 <0.001 -62 Stride length α 0.67 (0.12) 0.31 (0.12) 0.33 (0.15) 135 <0.001 0.79 0.73 0.86 <0.001 -54 <0.001 -51 Stride speed α 0.32 (0.11) 0.32 (0.11) 0.29 (0.15) 1.43 0.25 0.04 0.00 0.21 - - - - Thirty-six subjects walked on a treadmill under 3 conditions: NC = no cueing (normal walking); AC = auditory cueing (metronome walking); VC = visual cueing (visually guided stepping). In each condition for each subject, 500 gait cycles were recorded, from which the serial correlations of the spatiotemporal gait parameters were assessed (scaling exponent α, detrended fluctuation analysis). Descriptive and inferential statistics and significant post hoc comparisons are shown. % is the relative difference: (VC-NC)/NC *100. Significant results are shown in bold. Author's accepted version Rev2, Feb. 2016 Fractal fluctuations of guided stepping 21 TABLE 3. Overview of the cueing effects on fractal dynamics and fluctuations magnitude in human walking. Stride time Stride length Stride speed Overground walking16, 42 Overground walking + AC42 α >0.5 / fl α <0.5 / fl Overground walking + VC ? Treadmill walking8, 34, 41 Treadmill walking + AC34, 41 Treadmill walking + VC α >0.5 / fl α <0.5 / fl α <0.5 / › α >0.5 / fl α >0.5 / fl ? α >0.5 / fl α <0.5 / fl α <0.5 / › α >0.5 / fl α >0.5 / fl ? α <0.5 / fl α <0.5 / fl α <0.5 / › Each row shows the results for a particular cueing condition on the different gait parameters (columns). AC: auditory cueing. VC: visual cueing. α >0.5 (green): DFA indicates the presence of long- range correlations. α <0.5 (red): DFA indicates the presence of anti-correlations. fl fluctuation magnitude. › (red): high fluctuation magnitude. The new findings brought by the present (green): low study are shown in bold. The numbers refers to the bibliography at the end of the article. Author's accepted version Rev2, Feb. 2016 › › › › › › › › ›
1810.05077
1
1810
2018-10-10T09:22:21
On the Brain Networks of Complex Problem Solving
[ "q-bio.NC", "cs.CV" ]
Complex problem solving is a high level cognitive process which has been thoroughly studied over the last decade. The Tower of London (TOL) is a task that has been widely used to study problem-solving. In this study, we aim to explore the underlying cognitive network dynamics among anatomical regions of complex problem solving and its sub-phases, namely planning and execution. A new brain network construction model establishing dynamic functional brain networks using fMRI is proposed. The first step of the model is a preprocessing pipeline that manages to decrease the spatial redundancy while increasing the temporal resolution of the fMRI recordings. Then, dynamic brain networks are estimated using artificial neural networks. The network properties of the estimated brain networks are studied in order to identify regions of interest, such as hubs and subgroups of densely connected brain regions. The major similarities and dissimilarities of the network structure of planning and execution phases are highlighted. Our findings show the hubs and clusters of densely interconnected regions during both subtasks. It is observed that there are more hubs during the planning phase compared to the execution phase, and the clusters are more strongly connected during planning compared to execution.
q-bio.NC
q-bio
ON THE BRAIN NETWORKS OF COMPLEX PROBLEM SOLVING Abdullah Alchihabi † Omer Ekmekci † Baran B. Kivilcim † Sharlene D. Newman(cid:63) Fatos T. Yarman Vural † Middle East Technical University, {abdullah.alchihabi, oekmekci, baran.kivilcim, vural}@ceng.metu.edu.tr † Department of Computer Engineering (cid:63) Department of Psychological and Brain Sciences Indiana University , [email protected] ABSTRACT Complex problem solving is a high level cognitive process which has been thoroughly studied over the last decade. The Tower of London (TOL) is a task that has been widely used to study problem-solving. In this study, we aim to explore the underlying cognitive network dynamics among anatomi- cal regions of complex problem solving and its sub-phases, namely planning and execution. A new brain network con- struction model establishing dynamic functional brain net- works using fMRI is proposed. The first step of the model is a preprocessing pipeline that manages to decrease the spa- tial redundancy while increasing the temporal resolution of the fMRI recordings. Then, dynamic brain networks are es- timated using artificial neural networks. The network prop- erties of the estimated brain networks are studied in order to identify regions of interest, such as hubs and subgroups of densely connected brain regions. The major similarities and dissimilarities of the network structure of planning and execu- tion phases are highlighted. Our findings show the hubs and clusters of densely interconnected regions during both sub- tasks. It is observed that there are more hubs during the plan- ning phase compared to the execution phase, and the clusters are more strongly connected during planning compared to ex- ecution. Index Terms -- fMRI, Machine Learning, Brain Net- works, Tower of London, Complex Problem Solving. 1. INTRODUCTION Complex problem solving has been the focus of numerous studies in the field of neuroscience for over 30 years given the large number of high-level cognitive tasks that fall under its umbrella. To name a few, complex problem solving includes, strategy formation, coordination, sequencing of mental func- tions, and holding information on-line. These complex high- level cognitive sub-processes make revealing the inner work- ings of problem solving difficult. The Tower of London (TOL) task, designed by Tim Shal- lice in (1982) [1], has been one of the standard tools in the lit- erature to study complex problem solving. It consists of three bins having different capacities with colored balls placed in the bins; the aim is to rearrange the balls from their initial state to a predetermined goal state while moving one ball at a time and taking into consideration the limited capacity of each bin. TOL also has been used to investigate the effect of various clinical disorders on functions associated with the prefrontal cortex such as planning. It has been utilized to identify exec- utive dysfunction in children and adolescents suffering from epilepsy and seizures [2]. It also has been used to analyze the cognitive activation patterns of individuals suffering from de- pression [3]. Additionally, it has been employed to examine cognitive impairment in patient's diagnosed with Parkinson disease [4]. In another study, TOL along with fMRI have been employed to study the differences in the neural basis of plan- ning and executive function between first-episode schizophre- nia patients and healthy subjects [5]. Besides clinical disorders, TOL has also been employed to study the effect of various parameters on complex prob- lem solving performance in healthy subjects. The predictive power of working memory, inhibition, and fluid intelligence on TOL performance has been explored [6, 7]. Also, the ef- fect of physical activity and exercise, age, gender, and im- pairment in the executive function on planning and problem- solving ability and its underlying neural basis have all been studied [8, 9, 10, 11, 12, 13, 14]. TOL itself has many variations due to its large number of parameters such as goal hierarchy, demand for subgoal gen- eration, number of solution paths, and the existence of sub- optimal alternatives. Several studies have examined the ef- fect of the aforementioned structural parameters along with numerous other non-structural parameters including instruc- tions, experience, environment and problem-solving strategy on the performance of subjects when solving TOL puzzles, where performance is measured by pre-planning time and ac- curacy [15, 16, 17, 18, 19, 20, 21]. Despite the popularity of the TOL problem in the litera- ture and the wide range of problems covered in these stud- ies, relatively few researchers have explored the underlying network structures. In [22], the involvement of the parietal cortex, prefrontal cortex, basal ganglia and anterior cingu- late in complex problem solving was reported. The activa- tion patterns of the dorsolateral and rostrolateral subregions of the prefrontal cortex during planning has been examined [23]. The focus of another study has been the hemispheric differences in the pre-frontal cortex during planning and exe- cution as well as the contribution of the superior parietal re- gion to spatial working memory [24]. In addition, some work has been done to investigate the variance in the neural basis of planning between standard and expert subjects [25]. Other works have investigated the cognitive activation patterns dur- ing planning and execution subtasks [26, 27, 28, 29]. Given this brief literature review, a holistic understanding of the ac- tive anatomical brain regions, their respective roles and their interactions during complex problem-solving is an important, yet lacking research study. Numerous studies have proposed various computational models in order to build brain networks from fMRI measure- ments, both during cognitive tasks or during resting state. These studies represent a shift in the literature towards brain decoding algorithms that are based on the connectivity pat- terns in the brain motivated by the findings that these patterns provide more information about cognitive tasks than the iso- lated behavior of individual voxel groups or anatomical re- gions [30, 31, 32, 33, 34]. Some of these studies focused on the pairwise relation- ships between voxels or brain regions. For example, Pear- son correlation has been used in order to construct undirected functional connectivity graphs at different frequency resolu- tions in [35]. Also, pairwise correlations and mutual informa- tion have also been used in order build functional brain net- works in various studies aiming to investigate the network dif- ferences between patients with Schizophrenia or Alzheimers disease and healthy subjects [36, 37, 38]. Partial correlation along with constrained linear regression was also used to gen- erate brain networks in [39]. Other studies take advantage of the locality property of the brain by constructing local mesh networks around each brain region then representing the entire brain network as an ensem- ble of local meshes. In such studies, the Blood-Oxygenation Level Dependent (BOLD) response of each brain region is es- timated as a linear combination of the responses of its closest neighboring regions. Levinson-Durbin recursion has been ap- plied in several studies in order to estimate the edge weights of each local star mesh, where the nodes are the neighboring regions of the seed brain region [40, 41]. Ridge regression has also been used to estimate edge weights while construct- ing local mesh networks across windows of time [42, 32]. Other works suggested various methods to prune the constructed brain networks. In [43], researchers established resting-state brain networks as sparse constrained networks using both L1 and L2 regularization to introduce sparsity and control for across-subject variability. Another study used a two-step model to build functional brain networks, where at first a sparse multivariate autoregressive model was employed with penalized regression to estimate the brain networks. Then, false discovery rate (FDR) is used to prune low probability connections to present sparsity in the brain network [44]. After constructing brain networks, the statistical proper- ties of the established networks are studied in order to ob- tain neuroscientific insights related to the experiment at hand. Several measures of centrality have been proposed that aim to identify potential hubs that are central to the flow of infor- mation in the network such as: node degree, node strength and node betweenness centrality. In addition, measures of functional segregation have been proposed that aim to detect subgroups or modules of densely interconnected anatomic re- gions such as: local efficiency, clustering coefficient and tran- sitivity [45, 46, 47]. Furthermore, the node degree distribution of constructed brain networks has been measured and com- pared with power law and truncated power law distributions [48]. Also, the small-world property of brain networks has been studied extensively in numerous studies [48, 46]. Some studies, such as [49], have further extended the literature by defining the null model for weighted undirected functional brain networks. Further work have focused on controlling for family-wise error (FWE) that complements false discov- ery rate (FDR) [50]. Several studies have compared the network properties of functional brain networks across different age groups [51, 47] and under different developmental factors [52]. Other stud- ies have performed similar analysis to compare the network properties of healthy individuals against those suffering from several diseases related to cognitive impairment (Alzheimer, epilepsy, Schizophrenia) [53]. In this study, we propose a dynamic functional brain net- work model, extracted from fMRI measurements using arti- ficial neural networks. The decoding power of the suggested brain network model is examined by distinguishing the two phases of complex problem solving, namely, planning and execution. Then, the network properties of the dynamic brain networks are studied in order to identify the active anatomical regions during both planning and execution phases of com- plex problem solving. Potential hubs and clusters of densely connected brain regions are identified for both subtasks. Fur- thermore, the distinctions and similarities between planning and execution networks are highlighted. The results identify both active, inactive, hub regions as well as clusters of densely connected anatomical regions during complex problem task. In addition, results show that there are potential hubs during planning phase compared to execution phase, also the clus- ters of densely interconnected regions are significantly more strongly connected during planning compared to execution. 2. TOL EXPERIMENT PROCEDURE In this section, we introduce the details of the experiment as well as data collection and preprocessing methods. 2.1. Participants & Stimuli 18 college students aged between 19 and 38 participated in the experiment after signing informed, written consent doc- uments approved by the Indiana University Institutional Re- view Board. The subjects solved a computerized version of TOL problem, two configurations were presented at the be- ginning of each puzzle: the initial state and the goal state. The subjects were asked to transform the initial state into the goal state using the minimum number of moves. However, the subjects were not informed of the minimum number of moves needed to solve a given puzzle nor of the existence of multiple solution paths. 2.2. Procedure Each subject underwent a practice session before entering the scanning session to acquaint subjects with the TOL problem. The subjects were given the following instructions: You will be asked to solve a series of puzzles. The goal of the puz- zle is to make the start or current state match the goal state (They were shown an example). Try to solve the problems in the minimum number of moves by planning ahead. Work as quickly and accurately as possible, but accuracy is more important than speed. The scanning session consisted of 4 runs, each run in- cluded 18 timed puzzles, with a 5 second planning only time slot during which subjects were not allowed to move the balls. However, they were allowed to continue planning after the 5 seconds planning only time slot if they chose to do so. Fol- lowing every puzzle, there was a 12-second rest period where subjects focused on a plus sign in the center of the screen. Each run was also followed by a 28-second fixation period. 2.3. fMRI Data Acquisition & Preliminary Analysis The fMRI images were collected using a 3 T Siemens TRIO scanner with an 8-channel radio frequency coil located in the Imaging Research Facility at Indiana University. The images were acquired in 18 5 mm thick oblique axial slices using the following set of parameters: TR=1000 ms, TE=25 ms, flip angle=60◦, voxel size=3.125 mm×3.125 mm×5 mm with a 1 mm gap. The statistical parametric mapping toolbox was used to perform the preliminary data analysis that included: image correction for slice acquisition timing, resampling, spatial smoothing, motion correction and normalization to the Mon- treal Neurological Institute (MNI) EPI template. Further details concerning the procedure and data acquisition can be found in [22]. 3. MESH BRAIN NETWORKS FOR COMPLEX PROBLEM SOLVING 3.1. Data Processing Given the small number of subjects in the TOL dataset, and the large number of voxels in each brain volume (185,405 voxels per time instant), voxel selection is used to reduce the spatial resolution of the collected brain images and dampen the noise that is inherent in the data. Furthermore, due to the short duration of each puzzle (max 15 seconds) and the rela- tively low sampling rate (TR = 1 sec), temporal interpolation is needed in order to increase the number of brain volumes for each puzzle. Finally, Gaussian noise is used in order to regularize the data and improve generalization. 3.1.1. Voxel Selection First, an ANOVA feature selection method is used to choose the most discriminative subset of voxels and discard the re- maining ones [54, 55, 56]. For this purpose, we calculate the f-value score of each voxel vi as shown in equation 1: f scorei = M SB(vi, ylabel) M SW (vi, ylabel) (1) where ylabel is the label indicating the subtask (Planning or Execution). M SB(vi, ylabel) is the mean square value be- tween voxel i and the label vector ylabel which is calculated by equation 2 M SB(vi, ylabel) = SSB(vi, ylabel) dfbetween , (2) SSB(vi, ylabel) is the sum of squares between ylabel and vi, dfbetween is the number of groups minus one. M SW (vi, ylabel) is the mean square value within voxel i and the label vector ylabel and it is calculated by 3 M SW (vi, ylabel) = SSW (vi, ylabel) dfwithin (3) where, SSW (vi, ylabel) is the sum of squares within group and dfwithin is the degree of freedom within (total number of elements in vi and ylabel minus the number of groups). We order the voxels according to their f-value scores. Then, the distribution of f-value scores of all voxels is plot- ted in order to determine the appropriate number of voxels to retain. Voxel selection is applied to the voxels of all brain regions except the ones located in the cerebellum, which we exclude during network extraction. Voxel selection successfully manages to significantly re- duce the number of voxels in each brain volume, thus, making the space and time complexity of the analysis on the dataset feasible given the large total number of voxels in each brain volume, 185,405 voxels per time instant. It is also a neces- sary step for decreasing the curse of dimensionality problem for decoding the planning and execution phases. The BOLD response of the selected voxels is then aver- aged into their corresponding brain regions defined by the automated anatomical labeling (AAL) atlas [57] as shown in equation 4: (cid:80) rj = i∈ζ[j] vi ζ[j] (4) where rj is the BOLD response of region j, vi is the BOLD response of voxel i and ζ[j] is the set of selected vox- els located in region j. Averaging the selected voxels in their corresponding anatomical regions smooths the noise embed- ded in fMRI signal to a certain degree, and further reduces the dimensionality of each brain volume. As a result, each region is represented by a BOLD response, thus, enabling us to investigate the role and contribution of each region to the planning and execution phases of the problem solving task. 3.1.2. Interpolation It is well-known that in spite of its high spatial resolution, fMRI signal has very low temporal resolution compared to EEG signal. In this study, we interpolate the fMRI signal in order to compensate for this drawback and study the effect of interpolation on decoding planning and execution phases of TOL. In the TOL study subjects solved a puzzle in at most 15 seconds and the sampling rate, TR, is 1000 ms. Interpolation is used to increase the temporal resolution by estimating z extra brain volumes between each two consecutive measured brain volumes. As a result, the total number of available brain volumes for each puzzle becomes n+z∗(n−1), where n is the number of measured brain volumes of a given puzzle. We use the cubic spline interpolation function rather than linear inter- polation methods in order to prevent edge effects and smooth- ing out the spikes between the measured brain volumes [58]. In order to analyze the effect of time interpolation and to estimate an acceptable number of inserted brain volumes z, we compare the Fourier Transform of the fMRI signal com- puted before and after interpolation so that the frequency con- tent of the signal is not distorted by interpolation. The origi- nal single-sided amplitude of the signal and the one obtained after interpolation are compared in order to ensure that in- terpolation is preserving the smooth peaks of the data in the frequency domain [59, 60]. 3.1.3. Injecting Gaussian Noise When modeling a deterministic signal by a probabilistic method, adding noise to the signal decreases the estimation error in most of the practical applications. The final phase of preprocessing is adding a Gaussian noise to the interpolated time series of the BOLD response in each anatomical region. For this purpose, instead of just injecting white noise, a rather informed noise, colorful Gaussian noise, is added. In order to reflect the corresponding brain region's properties, for each sample the additive noise sample is generated from a Gaus- sian distribution having mean and variance of that anatomical region. This newly generated samples not only act like a natural regularizer to improve generalization performance of brain decoding, but, also help making local mesh estima- tion algorithms more stable when generating brain networks [61, 62]. Given a representative time series from a particular brain region, i represents the index of an anatomical re- gion. The new samples are generated with vector addition of noise while preserving the signal-to-noise ratio (SNR) as in rj = rj + τj, where τj is a noise vector sampled from N (αnoise µ(rj), βnoise σ2(rj)), αnoise and βnoise are the scaling factors which are set empirically, to optimize the decoding performance. 3.2. Building Dynamic Brain Network with Neural Net- works After applying the preprocessing pipeline, we construct dy- namic functional brain networks. In order to do that, we par- tition each time series, which represents an anatomical region into fixed-size windows, where each window, win(t), is cen- tered at the measured brain volume at time instance, t. The size of each window is W in Size = z + 1 brain volumes, where z is the number of interpolated brain volumes in each window. Equation 5 shows the time instances included in each window. (cid:34) t −(cid:106) z (cid:107) 2 (cid:109)(cid:35) (cid:108) z 2 win(t) = , .., t, .., t + (5) We construct a dynamic brain network, N (t) = (V, W (t)), for each time window win(t), where V is the set of nodes of the graph corresponding to the brain anatomical regions, while W (t) = {wt,j,i∀i, j ∈ V } is the directed weighted edges between the nodes of the graph within time window win(t). The nodes of the graph are the AAL defined brain regions[57], except for the regions located in the cerebellum. The nodes are then pruned using voxel selection, as some anatomical regions contribute no voxels at all and get deleted from the set of nodes of the graph V . Note that our aim is to label the BOLD responses mea- sured at each brain volume as it belongs to one of the two phases of complex problem solving, namely, planning and execution. For this purpose, we represent each brain volume measured at a time instant t by a network, which shows the relationship among the anatomical regions. This dynamic net- work representation will allow us to investigate the network properties of planning and execution subtasks. In this section, we describe how we estimate the weights of the edges, W (t), of the brain network, N (t), for each time instance, t, where we employ the method proposed in [63]. For each window win(t), we define the functional neigh- borhood matrix, Ωt. The entries of Ωt are binary, either 1 or 0, indicating if there is a connection between two regions or not. The size of the matrix is M xM, where M is the num- ber of brain anatomical regions. The functional neighborhood matrix contains no self-connections, thus, Ωt(i, i) = 0∀i ∈ [1, M ]. Also, the brain regions pruned by voxel selection contributing no voxels have no in/out connections, thus, the corresponding entries in Ωt are all zeros. The connectivity of each region to the rest of the regions is determined using Pearson correlation, as follows: first, for every region i, we measure the Pearson correlation between its BOLD response ri,t and the BOLD responses of all the other remaining re- gions as shown below: cor(ri,t, rj,t) = cov(ri,t, rj,t) σ(ri,t)σ(rj,t) , (6) where ri,t is the BOLD response of region i across time window win(t), cov(ri,t, rj,t) is the covariance between the corresponding BOLD responses of regions i and j. σ is the standard deviation of the BOLD response of a given region. Thus, the higher the Pearson correlation between two regions the closer they are to each other in the functional neighbor- hood system. Then, we select p of the regions with the highest correla- tion scores with region i. Thus, obtaining the neighborhood set ηp[i], which contains the p closest brain regions to region i. Finally, we define the Ωt(i, j) as the connectivity between the regions i and j, using the constructed neighborhood sets as follows: (cid:40) Ωt(i, j) = if j ∈ ηp[i] 1, 0, otherwise. (7) Note that each anatomical region is connected to its p closest functional neighbors. This approach forms a star mesh around each anatomical region. The ensemble of all of the local meshes creates a brain network at each time instance. Note, also, that Pearson correlation values are not used as the weights between two regions. They are just used to identify the nodes of each local mesh formed around an anatomical re- gion. The estimated brain network becomes sparser as p gets smaller. When p is set to the number of anatomic regions, M, the network becomes fully connected. The selection method for the degree of neighborhood, p, is explained in the next section. This approach of defining the connectivity matrix not only makes the network representation sparse for small p values, but, it also constructs a network which is connected in functionally closest regions, satisfying the locality property of the human brain. After having determined the edges of the brain graph us- ing the functional neighborhood matrix Ωt, all that is left is to estimate the weights of these edges at each local mesh. In or- der to do that, we represent the response of each region i (ri,t) as a linear combination of its closest p-functional neighbors as shown in equation 8, (cid:88) j∈ηp[i] ri,t = wt,j,irj,t + i,t. (8) In equation 8, ri,t represents the representative time se- ries of the BOLD response of region i within the time window win(t), wt,j,i is the estimated edge weight between node (re- gion) i and j at time instance t. ηp[i] is the p closest functional neighbors of region i. Onal et.al. [64] estimated the arc-weights for each mesh formed around region i for each time window win(t) by min- imizing the mean-squared error loss function using Ridge re- gression. In this approach, the mean-squared error loss func- tion is minimized with respect to wt,j,i, for each region, inde- pendent of the other regions, where the expectation is taken over the time-instances, in window win(t) as shown in equa- tion 9 . E[(i,t)2] = E[(ri,t − (cid:88) j∈ηp[i] wt,j,irj,t)2] + λwt,j,i2, (9) where λ is the L2 regularization parameter whose value is optimized using cross-validation. L2 regularization is used in order to improve the generalization of the con- structed mesh networks. Note that the estimated arc-weights, wt,j,i (cid:54)= wt,i,j. Therefore, the ensemble of meshes yields a directed brain network. In this work, we define an artificial neural network to es- timate the values of mesh arc-weights for all anatomical re- gions jointly in each time window, as proposed in [63]. In this method, we estimate the mesh arc-weights matrix W (t) = {wt,j,ij, i ∈ V } using a feedforward neural network. The architecture of this network consists of an input layer and an output layer, both containing M nodes corresponding to each brain region. The edges of this network are constructed us- ing the neighborhood matrix Ωt. There is an edge between node i of the output layer and node j from the input layer, if Ωt(i, j) = 1. The loss function of the suggested artificial neural net- work is given in equation 10, where W is the weight matrix of the entire neural network that corresponds to directed edge weights of the brain graph and Wi is the row of matrix W corresponding to region i: Loss(Outputi) = E[(i,t)2] + λW T i Wi wt,j,irj,t)2] + λW T i Wi. = E[(ri,t − (cid:88) j∈ηp[i] We train the aforementioned artificial neural network in order to obtain the weights of the brain network at each time instance t that minimize the loss function by applying a gra- dient descent optimization method as shown in equation 11, (10) t,j,i = w(κ−1) w(κ) t,j,i − αlearning ∂E[(i,t)2] ∂wt,j,i , (11) where w(κ) t,j,i is the weight of the edge from node j to node i at epoch (iteration) κ, αlearning is the learning rate. The number of epochs and learning rate used to train the network are optimized empirically using cross-validation. Finally, the weights of the above artificial neural network, computed for each win(t), correspond to the edge weights of the dynamic brain network, N (t) = (V, W (t)), at each time instant t. Thus, we refer to the brain networks using their window indices in order to obtain a set of dynamic brain networks T = {N (1), N (2), ...N (tot win)}, where N (t) is the brain network for time window win(t) and tot win is the total number of time windows. 3.3. Network Metrics for Analyzing Brain Networks In this section, we introduce some measures which we will use to investigate the network properties of each phase of the complex problem solving task, namely, planning and execu- tion, using the estimated dynamic brain functional networks. The connectivity patterns of anatomical regions are analyzed by the set of network measures, given below. Two separate sets of measures are used, namely, measures of centrality and segregation. Since our estimated brain networks are directed, we distinguish the incoming and outgoing edges in the net- work, while defining the measures. Recall that the suggested brain network N (t) = (V, W (t)) consist of a set of nodes, V , each of which corresponding to one of the M anatomical regions. W (t) is the dynamic edge weight matrix with the entries, wi,j, representing the weight of the edge from node i to node j. For the sake of simplicity, we omit the time dependency parameter t, since we compute the network properties at each time instant. Matrix A is the binarized version of W (t) matrix, where ai,j takes value 0 if (wi,j == 0) and takes value 1 otherwise. 3.3.1. Measures of Centrality Measures of centrality aim to identify brain regions that play a central role in the flow of information in the brain network, or nodes that can be identified as hubs. It is commonly mea- sured using node degree, node strength and node betweenness centrality, which are defined below. 3.3.2. Node Degree The degree of a node is the total number of its edges as shown in equation 12, where degreei is the degree of node i, V is the set of all nodes in the graph and ai,j is the edge between node i and node j. (cid:88) j∈V ai,j degreei = (12) In the case of a directed graph, we distinguish two dif- i and node out-degree ferent metrics: node in-degree degreein i metrics which are shown in equations 13 and 14 degreeout respectively where aj,i = 1, if there is a directed edge from node j to node i. (cid:88) (cid:88) j∈V j∈V aj,i ai,j degreein i = degreeout i = (13) (14) Node degree is a measure of centrality of the given nodes, where it aims to quantify the hub brain regions interacting with a large number of brain regions. Thus, a node with high degree indicates its central role in the network. 3.3.3. Node Strength Node strength is the sum of the weights of edges connected to a given node 15, where wi,j is the weight of the edge between node i and node j. (cid:88) j∈V wi,j strengthi = (15) Similar to node degree, node strength, also, distinguishes two metrics in the case of directed graphs, namely, node in- strength strengthin shown in equations 16 and 17 respectively, where wj,i is the weight of the edge from node j to node i. i and out-strength strengthout i (cid:88) j∈V wj,i strengthin i = (16) (cid:88) j∈V strengthout i = wi,j. (17) Node strength is a node centrality measure that is sim- ilar to node degree, which is used in the case of weighted graphs. Nodes with large strength values are tightly con- nected to other nodes in the network forming hub nodes. 3.3.4. Node Betweenness Centrality Betweenness centrality of node i is the fraction of the shortest paths in the network that pass through node i as shown in equation 18 betweennessi = 1 (M − 1)(M − 2) (cid:88) j,k∈V ρi j,k ρj,k , (18) where ρj,k is the number of shortest paths betweens nodes j and k, ρi j,k is the number of shortest paths between nodes j and k that pass through node i, nodes i, j and k are distinct nodes. Before measuring the betweenness centrality of a node, we need to change our perspective from connection weight matrix to connection length matrix since betweenness central- ity is a distance-based metric. In connection weights matrix, larger weights imply higher correlation and shorter distance while it is the opposite in the case of length matrix. Connec- tion length matrix is obtained by inverting the weights of the connection weight matrix. Then, the algorithm suggested in [65] is employed in order to calculate the node betweenness centrality for each anatomical region. Nodes with high betweenness centrality are expected to participate in many of the shortest paths of the networks. Thus, taking a crucial role in the information flow of the network. proposed in [66]. It is defined as the fraction of the neighbors of node i that are also neighbors of each other. Ci = [(dout i + din i )(dout i − 1) − 2(cid:80) χi i + din j∈V ai,jaj,i] . (19) where din i is the in-degree of node i and dout is the out- degree of node i. χi is the weighted geometric mean of trian- gles around node i that is calculated by equation 20. Recall that aj,i = 1 , if there is a directed edge from node j to node i and aj,i = 0, otherwise. i (cid:88) j,h∈V 1 2 χi = (wi,jwi,hwj,h)1/3. (20) The clustering coefficient of a node is the fraction of trian- gles around the node. It is defined as the fraction of the neigh- bors of the node that are also the neighbors of each other. 3.3.7. Transitivity Transitivity of a node is similar to its clustering coefficient. However, transitivity is normalized over all nodes while clus- ter coefficient for each node is normalized independently which makes clustering coefficient biased towards nodes with low degree. Transitivity can be expressed as the ratio of tri- angles to triplets in the network. It is calculated by equation 21 , as suggested in [66]: (cid:80) Ti = j − 1) − 2(cid:80) χi j + din j∈V [(dout j + din j )(dout , h∈V aj,hah,j] (21) 3.3.5. Measures of Segregation Measures of segregation aim to quantify the existence of sub- groups within brain networks, where the nodes are densely interconnected. These subgroups are commonly referred to as clusters or modules. The existence of such clusters in func- tional brain networks is a sign of interdependence among the nodes forming the cluster. Measures of segregation include clustering coefficient, transitivity and local efficiency. While global efficiency is a measure of functional integration repre- senting how easy it is for information to flow in the network. 3.3.6. Clustering Coefficient The clustering coefficient of a node i is the fraction of tri- angles around node i which is calculated by equation 19 as is the in-degree of node j and dout where din is the out-degree j of node j. χi is the weighted geometric mean of triangles around node i that is calculated by equation 20. Note that ah,jaj,h = 1 , if there exits an edge in both directions. j 3.3.8. Global & Local Efficiency The global efficiency of a brain network is a measure of its functional integration. It measures the degree of communica- tion among the anatomical regions. Thus, it is closely related to the small-world property of a network. Formally speak- ing, global efficiency is defined as the average of the inverse shortest path lengths between all pairs of nodes in the brain network. Equation 22 shows how to calculate the global effi- ciency of a brain network, where w i,j is the weighted shortest path length between two distinct nodes i and j [45]. (cid:88) i∈V 1 M Eglobal = (cid:80) i,j)−1 j∈V (w M − 1 (22) On the other hand, the local efficiency of a network is defined as the global efficiency calculated over the neighbor- hood of a single node. The local efficiency is, thus, a measure of segregation rather than functional integration as it is closely related to clustering coefficient. While global efficiency is calculated for the entire network, local efficiency is calculated for each node in the network [45]. 4. EXPERIMENTS & RESULTS In this section, we explore the validity of the suggested net- work model by applying it to the TOL dataset. First, we an- alyze the effect of the preprocessing step on the brain decod- ing performance of planning and execution phases of com- plex problem solving. Then, we investigate the validity of the dynamic functional brain network model proposed in this study. Finally, we analyze the network properties of the con- structed functional brain networks for planning and execution subtasks. 4.1. Voxel Selection At the first step of the proposed computational model, we dis- carded all of the voxels located in the cerebellum anatomical regions. Then, we calculated the f-score for each one of the remaining voxel and order the obtained f-scores of the vox- els. Following that, we plotted the ordered f-scores of the voxels in order to determine the appropriate number of voxels to retain. Figure 1 shows the ordered f-scores of the voxels averaged across all subjects. It can be observed from this fig- ure that a relatively small number of voxels is crucial for dis- criminating the subtasks of problem solving while the remain- ing voxels do not have significant information concerning the subtasks of problem solving. Based on the f-score distribu- tion shown in Figure 1, we kept the 10,000 voxels with the highest f-scores given the clear the elbow point whereas we discarded the remaining ones. After selecting the 10,000 voxels with the highest f- scores of each session, we computed the number of selected voxels contained in each one of the 90 anatomical regions. We also calculated the percentage of selected voxels to the total number of voxels located in each anatomical region. Figure 2a shows the average number of voxels contributed by each region across all subjects with its corresponding stan- dard deviation, Figure 2b shows the average percentage of voxels contributed by each region across all subjects with its corresponding standard deviation. It is clear from these figures that a large number of re- gions contribute little to no voxels, such as the amygdala, Fig. 1: Ordered f-scores of voxels. caudate, heschl gyrus, hippocampus, pallidum, putamen, tem- poral pole, superior temporal cortex, thalamus and parahip- pocampus. A small number of regions contribute a signifi- cantly large number of voxels (over 300 voxels each) during complex problem solving, such as occipital, precentral, pre- cuneus and parietal regions. Furthermore, Figure 2b ensures that there is no bias against tiny anatomical regions with small number of voxels by normalizing the number of voxels selected from each re- gion by its total number of voxels. Figure 2b clearly shows that in the left prefrontal and inferior occipital regions a significant percentage of voxels are active during complex problem solving. Both figures also show high standard de- viations across subjects, which indicates high inter-subject variability. 4.2. Interpolation After selecting the most discriminative voxels and averag- ing their BOLD responses with respect to their corresponding brain anatomical regions, we employed temporal interpola- tion to increase the temporal resolution of the TOL dataset. As a result, the total number of obtained brain volumes is equal to n + z ∗ (n − 1) where n is the number of mea- sured brain volumes of a given puzzle and z is the number of estimated brain volumes plugged between each pair of mea- sured brain volumes. The optimal value of z is equal to 8 which is determined empirically using cross-validation. Fig- ure 3 shows the interpolated BOLD response of a randomly selected anatomical region from the given subjects, where the blue dots represent the measured BOLD response of the re- gion and the orange dashes are the interpolated values. It is clear from Figure 3 that the interpolated points using cu- bic spline function do not introduce sharp edges nor do they smooth out the spikes between measured brain volumes. Furthermore, Figure 4 shows the single-sided amplitude spectrum of a randomly selected anatomical region from a given subject before interpolation, after interpolation and fi- nally after adding Gaussian noise. The figure clearly demon- (a) Average number of voxels selected from each anatomical region across all subjects. (b) Average percentage of voxels selected from each anatomical region across all subjects. Fig. 2: Distribution of selected voxels across anatomical regions, measured by number of selected voxels (top) and percentage of selected voxels (bottom) from each anatomical region. [0.025, 0.05, 0.075, 0.1]. 4.4. Brain Decoding We use brain decoding in order to quantify the effect of our proposed preprocessing steps on the TOL dataset. We aim to distinguish the two phases of complex problem solving namely: planning and execution. At first, we used ANOVA to select the 10,000 voxels with the highest f-scores then we averaged the selected voxels into their corresponding anatom- ical regions defined by AAL [57]. Following that, we em- ployed temporal interpolation to increase the temporal res- olution of each puzzle by estimating z = 8 brain volumes between each pair of measured brain volumes. Finally, we added Gaussian noise in order to regularize the BOLD re- sponses of each region to improve the generalization perfor- mance of the classifiers. We used k-fold Cross validation for each subject in all of the experiments introduced in this sec- tion, with k = 8. After we obtained the results, we averaged them across the different fold, then we calculated the average and standard deviation across all subjects. We used both su- pervised and unsupervised brain decoding methods, a linear support-vector machine (SVM) [67] was used for supervised brain decoding while k-means clustering was used for unsu- pervised brain decoding. Table 1 shows the effect of our preprocessing pipeline on the brain decoding of complex problem solving subtasks. The first row shows the performances of brain decoding on the raw dataset without any preprocessing, simply averaging all of the Fig. 3: Interpolated BOLD response. strates that both interpolation and injecting Gaussian noise preserve the smooth peaks of the signal in the frequency do- main. 4.3. Gaussian Noise In order to control the signal-to-noise ration (SNR), we used cross-validation to choose the optimal pair of values for αnoise and βnoise, the ratios of mean and standard devia- tion of the added noise respectively. As a result, the op- timal values obtained are αnoise = 0.025 and βnoise = 0.075 from the following set of values αnoise, βnoise ∈ Fig. 4: Single-Sided amplitude spectrum. voxels into their corresponding anatomical regions. While the second row shows the results of applying voxel selection then averaging the selected voxels into their anatomical regions. The third row shows the results of brain decoding after ap- plying temporal interpolation, while the forth row shows the results after injecting the data with Gaussian noise. Preprocessing Raw data Voxel Selection Interpolation Noise addition SVM 0.60 ± 0.11 0.74 ± 0.12 0.81 ± 0.08 0.82 ± 0.08 k-Means 0.63 ± 0.09 0.85 ± 0.06 0.84 ± 0.06 0.85 ± 0.06 Table 1: Decoding performances of preprocessing pipeline after each step. From the results of the preprocessing experiments, it is observed that voxel selection improves the brain decoding performance for both supervised and unsupervised methods from %60 to %74 and from %63 to %85 respectively. This can be attributed to voxel selection retaining only the most discriminative voxels and trashing the remaining less infor- mative ones. In addition, voxel selection manages to sparsify the representation of the data since some brain regions con- tribute no voxels at all thus have a flat BOLD response. The table also shows that temporal interpolation further improves the supervised brain decoding performance from %74 to %81, this significant increase is due to increasing the number of brain volumes thus increasing the number of train- ing samples for the classifier. However, temporal interpola- tion slightly reduces the performance of unsupervised meth- ods from %85 to %84 which can be partially attributed to the estimated brain volumes during the transitions between the two phases of problem solving, planning and execution which reduces the separation between the two natural sub- groups. This is due to the method used to label the estimated brain volumes, where each estimated brain volume is given the labels of its closest neighboring measured brain volume. Finally, the addition of Gaussian noise slightly boosts the performance of both supervised and unsupervised methods from %81 to %82 and from %84 to %85 respectively. The ta- ble also shows high standard deviation across subjects, which is consistent with voxel selection plots, revealing high inter- subject variability. 4.5. Building Brain Networks In this section, we compare our model for building dynamic functional brain networks with some of the popular methods proposed in the literature in terms of their brain decoding power. Brain decoding can verify whether the constructed brain networks are good representatives of the underlying cognitive subtasks or not. For this purpose, we built brain networks as explained in the previous sections after having successfully applied the preprocessing pipeline. The optimal values for learning rate αlearning and number of epochs were chosen empirically us- ing cross-validation obtaining the following values respec- tively 1 ∗ 10−8 and 10. As for p, the number of neighbors used to represent each anatomical region, we chose p equal to the total number of regions which is 90, in this way, a fully- connected brain network is obtained at each time window. However, the total number of nodes is less than 90 given that some regions have flat BOLD responses therefore they were pruned along with all their edges from the brain network. We also constructed brain networks using Pearson corre- lation and ridge regression as proposed in [35, 64] and [32, 42] respectively in order to compare the performance of our methods with other works in the literature. In the case of Pearson correlation, the functional brain networks were con- structed using Pearson correlation scores between each pair of brain regions [35, 64]. As for the case of ridge regression, the mesh arc-weight descriptors were estimated using ridge regression in order to represent each region as a linear combi- nation of its neighbors [32, 42]. Table 2 shows the brain decoding results of the aforemen- tioned brain network construction methods compared against the results of multi-voxel pattern analysis (MVPA). The first row shows the brain decoding results of MVPA, while the sec- ond and third rows show the results of Pearson correlation and ridge regression methods respectively. The last row shows the brain decoding results of our proposed neural network model. The table clearly shows that both Pearson correlation and ridge regression fail to construct valid brain networks that are good representatives of the underlying cognitive tasks. However, our model managed to get brain decoding results similar or slightly better than those obtained from MVPA both in the cases of supervised and unsupervised methods. This can be attributed to the challenging nature of the TOL dataset, Pearson correlation does not manage to capture the interde- pendencies between the anatomical regions over short time windows. While ridge regression fails to correctly estimate the mesh arc-weights as it estimates the arc-weights for each region independently of the other ones. Our proposed model, with a relatively small number of epochs manages to obtain mesh arc-weight values that capture the activation patterns of anatomical regions and their relationships. Algorithm MVPA Pearson Ridge Regression Neural Networks SVM 0.82 ± 0.08 0.58 ± 0.05 0.56 ± 0.05 0.82 ± 0.10 k-Means 0.85 ± 0.06 0.57 ± 0.04 0.55 ± 0.02 0.87 ± 0.06 Table 2: Results of proposed model. 5. BRAIN NETWORK PROPERTIES In this section, we aim to analyze the network properties of the constructed functional brain networks. We investigate the network properties for each anatomical brain region during both planning and execution subtasks in order to understand which regions are most active and which regions work to- gether during each one of the two subtasks of complex prob- lem solving. Given that the constructed brain functional networks are both weighted, directed, fully-connected and contain both negative and positive weights, we preprocessed the networks before measuring their network properties. Firstly, we got rid of all the negative weights by shifting all the mesh arc-weights values by a positive quantity equal to the absolute value of the largest negative arc-weight. Then, we normalized the mesh arc-weights to ensure that all of weights are within the range of [0, 1]. Finally, we measured the network properties on the pruned brain graph, where the brain regions (nodes) contributing no voxels (have a flat BOLD response) and all of their corresponding arc-weights (edges) were deleted from the brain graph. Thus, the networks contained less than 90 regions with their corresponding edges. We used brain connectivity toolbox to calculate the investigated network properties [45]. In order to measure for centrality, the number of neigh- bors for each anatomical region (P) was chosen to be equal to 89, which is equal to the total number of neighbors for any given node as the total number of brain anatomical regions defined by the AAL atlas [57] after deleting the regions resid- ing in the cerebellum equals 90. In addition, since we pruned the nodes that correspond to regions from which no voxels were selected, our constructed brain networks were weighted directed fully-connected networks. Therefore, the in-degree, out-degree and total degree of all nodes in the graph were equal to the total number of anatomical regions retained after voxel selection. Therefore, we used node strength and node betweenness centrality to identify nodes with high centrality which are po- tential hubs in the brain networks controlling the flow of in- formation in the network. In our proposed model, the node in-strength of node i is the sum of the mesh arc-weight values which is estimated using our proposed neural network method in order to minimize the reconstruction error of the BOLD re- sponse of anatomical region i using its neighbors. Thus, node in-strength is not used as part of our network properties analy- ses, we rather used node out-strength to measure the centrality of all anatomical regions. As for measures of segregation, quantifying the existence of subgroups within brain networks is based on densely in- terconnected nodes. These subgroups are commonly referred to as clusters or modules. The existence of such clusters in functional brain networks is a sign of interdependence among the nodes forming the cluster. Therefore, clustering coeffi- cient, transitivity and local efficiency were measured in order to identify potential clusters with dense interconnections in the brain networks. 5.1. Planning & Execution Brain Networks In this section, we discuss the network properties of the plan- ning and execution networks. For each aforementioned net- work metric, we ranked the brain regions in descending order according to their score on that network measure for all sub- jects across all sessions. Then, we retained the 10 anatomi- cal regions with the highest scores. Following that, we mea- sured the frequency of occurrence of each brain region among the top 10 regions across all sessions in order to identify the shared regions and patterns across all subjects for both plan- ning and execution subtasks. The results of the analysis are shown in tables: table 3 shows the brain regions that have high scores for the reported network properties during plan- ning subtask, and table 4 shows the brain regions that have high scores during execution subtask. There are a number of processes taking place during plan- ning and execution. Plan generation involves a series of re- cursive events including: 1) problem encoding; 2) decision- making in order to decide which ball to move and where to move it; 3) mental imagery to imagine the ball moving; and 4) working memory to maintain the intermediate steps as well as the move number. During plan execution there is 1) retrieval of the steps from memory; 2) confirming the correct steps are being performed; and 3) the motor execution of those steps. As the results demonstrate the networks for planning and ex- ecution are overlapping. These results are similar to the ac- tivation results reported in [22] in that the regions that were found to be active during the task are also regions that are most prominently found with the highest network measures. These regions include the right and left middle frontal gyrus, anterior cingulate cortex, precentral cortex, and superior pari- etal cortex. Previous work has suggested that the regions found in the current study to show high network measures are directly re- lated to the sub-tasks associated with TOL performance. For example, both the left and right prefrontal cortex have been found to be involved in the TOL task with the two regions per- forming distinguishable functions. The right prefrontal cortex is involved in constructing the plan for solving the TOL prob- lem while the left prefrontal cortex is involved in supervising the execution of that plan [22, 24]. The anterior cingulate has been linked to error detection and is particularly involved in the TOL when the number of moves is higher or the problem difficulty is manipulated. The right superior parietal cortex and precentral cortex have been linked to visuo-spatial atten- tion necessary for planning [24] and the left parietal cortex has been linked to visuo-spatial working memory processing [24]. The overlap between the regions with the highest net- work measures and those that have been linked to the task is an important feature and is not due to the voxel selection pro- cess. Many regions that passed threshold were not in the top ranked list of network measures. For example, the basal gan- glia including the caudate has been found in previous studies to be involved in TOL performance [22, 29, 68, 69, 26]; how- ever, the region appears to not be an important network hub. Figures 5a and 5b visualize the reported brain regions in tables 3 and 4 respectively using Brain Net Viewer [70]. In figures 5a and 5b, the color of the node (brain region) implies the following: red indicates that the region has high transi- tivity, clustering coefficient or local efficiency. Green indi- cates that the node has high node centrality measured by node out-strength and node betweenness. As for blue, it shows the nodes that have high node centrality and is part of subgroup of densely interconnected regions. 5.2. Differences between Planning and Execution Net- works In this section, we explore the network differences between planning and execution by calculating the difference between the network property scores for planning and execution for each session. To achieve that, we took the difference be- tween the network property scores for brain anatomical re- gions during planning and the network property scores for brain anatomical regions during execution for each session. Then, we counted the frequency of times a given anatomi- cal region is more active during planning than execution and vice-versa in order to identify consistent patterns of the dis- agreements between planning brain networks and execution brain networks across all subjects. Results showed, generally, that the network measures were higher for planning than ex- ecution. This, too, mirrors the findings from [22] in which planning resulted in greater activation than execution. Node out-strength is a measure of how connected the node is to other nodes in the network. Planning showed greater out- strength than execution in the following regions: occipital regions (calcarine, cuneus), parietal regions (bilateral supe- rior parietal cortex and precunues), the right superior frontal cortex, and inferior occipito-temporal regions (fusiform and lingual gyri). The left angular gyrus and bilateral medial su- perior frontal cortex showed greater out-strength for execu- tion. As for node betweenness, the following brain regions had higher node betweenness during planning than execution: occipital regions (calcarine, cuneus, right middle, right supe- rior); inferior occipito-temporal (fusiform, lingual); parietal (bilateral superior parietal, left postcentral, precuneus). Bi- lateral medial superior frontal had higher node betweenness during execution than planning. These results suggest that there is greater information flow during planning than execution. This matches our expecta- tions. Planning is more computationally demanding than ex- ecution. Again, during planning participants must explore the problem space which requires generating and manipu- transitivity Angular Calcarine Cingulum Ant Cingulum Mid Cuneus Frontal Inf Oper L local efficiency clustering coefficient Calcarine Cuneus Frontal Mid R Frontal Sup Fusiform Occipital Inf R Precentral Supp Motor Area R Temporal Inf R Calcarine Cuneus Frontal Mid R Frontal Sup Fusiform Occipital Inf R Parietal Sup R Precentral Supp Motor Area R Temporal Inf R betweenness Cueneus R Frontal Sup L Fusiform R Paracentral Lobule L Parietal Sup R Precuneus L Supp Motor Area R Temporal Inf R Temporal Mid R out-strength Cueneus R Frontal Sup L Fusiform R Paracentral Lobule L Supp Motor Area R Temporal Inf R Temporal Mid R Table 3: Planning: Anatomical regions with the highest network measures across subjects, regions are painted if they overlap with execution. transitivity Angular Calcarine Cingulum Ant Cingulum Mid Cuneus local efficiency Calcarine L Cuneus Frontal Sup L Fusiform Occipital Inf R Frontal Inf Oper Supp Motor Area R Temporal Inf R clustering coefficient Calcarine L Cuneus Frontal Mid R Frontal Sup Fusiform Occipital Inf R Parietal Sup R Precentral R Supp Motor Area R Temporal Inf R betweenness Cueneus R Frontal Sup L Fusiform R Paracentral Lobule L Precuneus L Supp Motor Area R Temporal Inf R Temporal Mid R out-strength Frontal Sup Fusiform Paracentral Lobule L Supp Motor Area R Temporal Inf R Temporal Mid R Table 4: Execution: Anatomical regions with the highest network measures across subjects, regions are painted if they overlap with planning. lating a mental representation of the problem. The regions that show greater information flow during planning are all regions involved in that generation and manipulation partic- ularly parietal, occipital and inferior occipito-temporal. On the other hand, execution requires recall of the plan gener- ated and stored and therefore, greater information flow from frontal regions related to memory retrieval is observed. Clustering coefficient, local efficency and transitvity are measures of segregation which aim to identify sub-networks. Each of these measures were larger for planning than exe- cution with no regions showing larger measures for execu- tion. The regions that showed higher clustering coefficient in planning included: the cuneus, left middle occipital cor- tex, and right precuneus. Local efficiency was higher in a similar set of regions (the cuneus, left middle occipital cor- tex, and right precuneus). The clustering coeffiencent and lo- cal efficiency identified a visual-spatial sub-network that is more strongly connected during planning. Transitivity iden- tified an overlapping but more extensive set of regions that included: bilateral angular gyrus, calcarine sulcus, cuneus, bilateral middle frontal cortex, bilaterial superior frontal cor- tex, bilateral fusiform and lingual gyri, bilateral occipital cor- tex, bilatral superior parietal cortex, postcentral and precen- tral cortex, precuneus, supplementary motor area, right supra- marginal gyrus, and right inferior and middle temporal cortex. Figures 6a , 6b visualize the brain regions with higher be- tweenness during planning and during execution respectively. Figures 7a , 7b visualize the brain regions with higher node out-strength during planning and during execution, re- spectively. Figure 8a visualizes the brain regions with higher local efficiency and higher clustering coefficient during planning phase compared to execution phase. While Figure 8b visual- izes the brain regions with higher transitivity during planning than during execution phase. 5.3. Global Efficiency Since global efficiency is measured over the entire brain net- work, not for a given node in the network, we measured the global efficiency for all planning and execution networks within all sessions across subjects. Then, global efficiency of planning is compared against that of execution. Results show that the majority of sessions had higher global efficiency scores during planning than execution, 43 out of 72 sessions had higher global efficiency during planning than execution. Furthermore, table 5 shows the number of sessions where global efficiency was higher during planning and during ex- ecution across all subjects for all 4 sessions of each subject. The first column shows the number of subjects that had a higher global efficiency score during planning than during execution. The second column shows the number of subjects that had a higher global efficiency score during execution than during planning. Although there was no significant difference in global efficiency between planning and execution, from the table, it is clear that the majority of subjects had a higher global efficiency for planning for the first session. Some subjects switched from having higher global efficiency during plan- ning to having higher global efficiency during execution. A potential explanation for this change across sessions is a switch from pre-planning to on-line planning, or planning intermixed with execution. Although there is a dedicated planning phase in the current study, that does not mean that planning is not taking place during execution. In fact, it has been debated as to whether efficient pre-planning is possible in the TOL or whether TOL performance is controlled by on-line planning [71, 72, 73, 6]. According to Phillips et.al. [71, 72] pre-planning the entire sequence is not natural, but that people instead plan the beginning sequence of moves and then intersperse planning and execution. If this is the case then it may be expected that some participants will switch to on-line planning. This intermixing of planning and execu- tion is also likely to impact the performance of the machine learning algorithms to detect planning and execution phases. The relationship between global efficiency and behavioral performance was examined. Global efficiency was found to be positively correlated with the mean number of extra moves (a measure of error) during problem-solving (for execution r=0.73, p=0.0006). Previous studies have shown a relation- ship between global efficiency and task performance [74]. This suggests that the variance in global efficiency is in- dicative of individual differences in neural processing and fur- ther suggests that the changes in global efficiency across ses- sions are also likely indicative of changes in neural processing related to changing strategy. Further research using a larger sample is necessary to explore this hypothesis. Session Number Planning Execution 1 2 3 4 15 9 10 9 3 9 8 9 Table 5: Global Efficiency. 6. CONCLUSION In this paper, we proposed a model to construct brain func- tional networks during a complex problem solving task. Our model successfully identified the two phases of complex problem solving. In addition, the network properties of the constructed brain networks during planning and execution phases were studied in order to identify essential nodes within the brain networks related to problem solving, potential hubs, and densely connected clusters. Furthermore, the differences between planning networks and execution networks were highlighted and discussed. There are some limitations to the study. Although the primary aim of this study was to demonstrate the feasibility of the methods, the sample size is somewhat small, making the interpretation of the results difficult. Second, a goal of this method is to identify brain states that are interspersed with each other. In the current study planning was expected to occur both prior to execution as well as during execution therefore planning states are interspersed within the execution phase. The temporal sampling rate of the fMRI data may be a limiting factor. Alternatively, the sluggish and blurred un- derlying hemodynamic response may be the factor preventing the ability to detect brain states. We plan to explore this factor in future work. 7. ACKNOWLEDGMENT The work is supported by TUBITAK (Scientific and Tech- nological Research Council of Turkey) under the grant No: 116E091 as well as the Indiana METACyt Initiative of Indi- ana University, funded in part through a major grant from the Lilly Endowment, Inc. 8. REFERENCES [1] Tim Shallice, "Specific impairments of planning," Phil. Trans. R. Soc. Lond. B, vol. 298, no. 1089, pp. 199 -- 209, 1982. [2] William S MacAllister, H Allison Bender, Lind- say Whitman, Antoinette Welsh, Shari Keller, Yael Granader, and Elisabeth MS Sherman, "Assessment of executive functioning in childhood epilepsy: the tower of london and brief," Child Neuropsychology, vol. 18, no. 4, pp. 404 -- 415, 2012. [3] Ingeborg Goethals, Kurt Audenaert, Filip Jacobs, Christophe Van de Wiele, Hamphrey Ham, Hanneke Pyck, Andr´e Vandierendonck, Cees Van Heeringen, and Rudi Dierckx, "Blunted prefrontal perfusion in de- pressed patients performing the tower of london task," Psychiatry Research: Neuroimaging, vol. 139, no. 1, pp. 31 -- 40, 2005. [4] Irena Rektorova, Hana Srovnalova, Radka Kubikova, and Jiri Prasek, "Striatal dopamine transporter imag- ing correlates with depressive symptoms and tower of london task performance in parkinson's disease," Move- ment disorders, vol. 23, no. 11, pp. 1580 -- 1587, 2008. [5] Paul E Rasser, Patrick Johnston, Jim Lagopoulos, Philip B Ward, Ulrich Schall, Renate Thienel, Stefan Bender, Arthur W Toga, and Paul M Thompson, "Func- tional mri bold response to tower of london performance of first-episode schizophrenia patients using cortical pat- tern matching," Neuroimage, vol. 26, no. 3, pp. 941 -- 951, 2005. [6] Josef M Unterrainer, Benjamin Rahm, Christoph P Kaller, Rainer Leonhart, K Quiske, K Hoppe-Seyler, C Meier, C Muller, and U Halsband, "Planning abili- ties and the tower of london: is this task measuring a discrete cognitive function?," Journal of clinical and experimental neuropsychology, vol. 26, no. 6, pp. 846 -- 856, 2004. [7] Nancy A Zook, Deana B Davalos, Edward L DeLosh, and Hasker P Davis, "Working memory, inhibition, and fluid intelligence as predictors of performance on tower of hanoi and london tasks," Brain and cognition, vol. 56, no. 3, pp. 286 -- 292, 2004. [8] Yu-Kai Chang, Chia-Liang Tsai, Tsung-Min Hung, Ed- mund Cheung So, Feng-Tzu Chen, and Jennifer L Et- nier, "Effects of acute exercise on executive function: a study with a tower of london task," Journal of Sport and Exercise Psychology, vol. 33, no. 6, pp. 847 -- 865, 2011. [9] Dustin Albert and Laurence Steinberg, "Age differences in strategic planning as indexed by the tower of london," Child development, vol. 82, no. 5, pp. 1501 -- 1517, 2011. [10] Nancy Zook, Marilyn C Welsh, and Vanessa Ewing, "Performance of healthy, older adults on the tower of london revised: Associations with verbal and nonverbal abilities," Aging, Neuropsychology, and Cognition, vol. 13, no. 1, pp. 1 -- 19, 2006. [11] A Boghi, R Rasetti, F Avidano, C Manzone, L Orsi, F D'agata, P Caroppo, M Bergui, P Rocca, L Pulvirenti, et al., "The effect of gender on planning: An fmri study using the tower of london task," Neuroimage, vol. 33, no. 3, pp. 999 -- 1010, 2006. [12] Josef M Unterrainer, Christian C Ruff, Benjamin Rahm, Christoph P Kaller, J Spreer, Ralf Schwarzwald, and Ul- rike Halsband, "The influence of sex differences and individual task performance on brain activation during planning," Neuroimage, vol. 24, no. 2, pp. 586 -- 590, 2005. [13] Christoph P Kaller, Katharina Heinze, Irina Mader, Josef M Unterrainer, Benjamin Rahm, Cornelius Weiller, and Lena Kostering, "Linking planning per- formance and gray matter density in mid-dorsolateral prefrontal cortex: moderating effects of age and sex," Neuroimage, vol. 63, no. 3, pp. 1454 -- 1463, 2012. [14] Manuel Desco, Francisco J Navas-Sanchez, Javier Sanchez-Gonz´alez, Santiago Reig, Olalla Robles, Car- Juan A Guzm´an-De-Villoria, Pedro olina Franco, Garc´ıa-Barreno, and Celso Arango, "Mathematically gifted adolescents use more extensive and more bilat- eral areas of the fronto-parietal network than controls during executive functioning and fluid reasoning tasks," Neuroimage, vol. 57, no. 1, pp. 281 -- 292, 2011. [15] Christoph P Kaller, Josef M Unterrainer, Benjamin Rahm, and Ulrike Halsband, "The impact of problem structure on planning: Insights from the tower of lon- don task," Cognitive Brain Research, vol. 20, no. 3, pp. 462 -- 472, 2004. [16] W Keith Berg, Dana L Byrd, Joseph PH McNamara, and Kimberly Case, "Deconstructing the tower: Param- eters and predictors of problem difficulty on the tower of london task," Brain and Cognition, vol. 72, no. 3, pp. 472 -- 482, 2010. [17] JM Unterrainer, B Rahm, U Halsband, and CP Kaller, "What is in a name: Comparing the tower of london with one of its variants," Cognitive Brain Research, vol. 23, no. 2, pp. 418 -- 428, 2005. [18] Sharlene D Newman and Gregory Pittman, "The tower of london: A study of the effect of problem structure on planning," Journal of Clinical and Experimental Neu- ropsychology, vol. 29, no. 3, pp. 333 -- 342, 2007. [19] JM Unterrainer, B Rahm, R Leonhart, CC Ruff, and U Halsband, "The tower of london: The impact of in- structions, cueing, and learning on planning abilities," Cognitive Brain Research, vol. 17, no. 3, pp. 675 -- 683, 2003. [20] Zachariah Campbell, Konstantine K Zakzanis, Diana Jo- vanovski, Steve Joordens, Richard Mraz, and Simon J Graham, "Utilizing virtual reality to improve the eco- logical validity of clinical neuropsychology: an fmri case study elucidating the neural basis of planning by comparing the tower of london with a three-dimensional navigation task," Applied Neuropsychology, vol. 16, no. 4, pp. 295 -- 306, 2009. [21] Sharlene D Newman, Benjamin Pruce, Akash Rusia, and Thomas Burns Jr, "The effect of strategy on prob- lem solving: an fmri study," The Journal of Problem Solving, vol. 3, no. 1, pp. 2, 2010. [22] Sharlene D Newman, John A Greco, and Donghoon Lee, "An fmri study of the tower of london: a look at problem structure differences," Brain research, vol. 1286, pp. 123 -- 132, 2009. [23] Gerd Wagner, Kathrin Koch, Jurgen R Reichenbach, Heinrich Sauer, and Ralf GM Schlosser, "The special in- volvement of the rostrolateral prefrontal cortex in plan- ning abilities: an event-related fmri study with the tower of london paradigm," Neuropsychologia, vol. 44, no. 12, pp. 2337 -- 2347, 2006. [24] Sharlene D Newman, Patricia A Carpenter, Sashank Varma, and Marcel Adam Just, "Frontal and parietal participation in problem solving in the tower of london: fmri and computational modeling of planning and high- level perception," Neuropsychologia, vol. 41, no. 12, pp. 1668 -- 1682, 2003. [25] F Cazalis, R Valabregue, M P´el´egrini-Issac, S Asloun, TW Robbins, and S Granon, "Individual differences in prefrontal cortical activation on the tower of london planning task: implication for effortful processing," Eu- ropean Journal of Neuroscience, vol. 17, no. 10, pp. 2219 -- 2225, 2003. [26] Odile A van den Heuvel, Henk J Groenewegen, Fred- erik Barkhof, Richard HC Lazeron, Richard van Dyck, and Dick J Veltman, "Frontostriatal system in planning complexity: a parametric functional magnetic resonance version of tower of london task," Neuroimage, vol. 18, no. 2, pp. 367 -- 374, 2003. [27] R Nathan Spreng, W Dale Stevens, Jon P Chamberlain, Adrian W Gilmore, and Daniel L Schacter, "Default network activity, coupled with the frontoparietal control network, supports goal-directed cognition," Neuroim- age, vol. 53, no. 1, pp. 303 -- 317, 2010. [28] Judy A Kipping, Daniel S Margulies, Simon B Eick- hoff, Annie Lee, and Anqi Qiu, "Trade-off of cerebello- cortical and cortico-cortical functional networks for planning in 6-year-old children," NeuroImage, vol. 176, pp. 510 -- 517, 2018. [29] MH Beauchamp, A Dagher, JAD Aston, and J Doyon, "Dynamic functional changes associated with cognitive skill learning of an adapted version of the tower of lon- don task," Neuroimage, vol. 20, no. 3, pp. 1649 -- 1660, 2003. [30] WR Shirer, S Ryali, E Rykhlevskaia, V Menon, and MD Greicius, "Decoding Subject-Driven Cognitive States with Whole-Brain Connectivity Patterns," Cere- bral Cortex, vol. 22, no. 1, pp. 158 -- 165, 2012. [31] Matthias Ekman, Jan Derrfuss, Marc Tittgemeyer, and Christian J Fiebach, "Predicting Errors from Reconfigu- ration Patterns in Human Brain Networks," Proceedings of the National Academy of Sciences, vol. 109, no. 41, pp. 16714 -- 16719, 2012. [32] Itir Onal, Mete Ozay, Eda Mizrak, Ilke Oztekin, and Fatos T Yarman Vural, "A new representation of fmri signal by a set of local meshes for brain decoding," IEEE Transactions on Signal and Information Processing over Networks, vol. 3, no. 4, pp. 683 -- 694, 2017. [33] Martin A Lindquist, "The Statistical Analysis of fMRI Data," Statistical Science, pp. 439 -- 464, 2008. [34] Jonas Richiardi, Sophie Achard, Horst Bunke, and Dim- itri Van De Ville, "Machine Learning with Brain Graphs: Predictive Modeling Approaches for Func- tional Imaging in Systems Neuroscience," IEEE Signal Processing Magazine, vol. 30, no. 3, pp. 58 -- 70, 2013. [35] Jonas Richiardi, Hamdi Eryilmaz, Sophie Schwartz, Pa- trik Vuilleumier, and Dimitri Van De Ville, "Decoding brain states from fmri connectivity graphs," Neuroim- age, vol. 56, no. 2, pp. 616 -- 626, 2011. [36] Mary-Ellen Lynall, Danielle S Bassett, Robert Kerwin, Peter J McKenna, Manfred Kitzbichler, Ulrich Muller, and Ed Bullmore, "Functional connectivity and brain networks in schizophrenia," Journal of Neuroscience, vol. 30, no. 28, pp. 9477 -- 9487, 2010. [37] Vinod Menon, "Large-scale brain networks and psy- chopathology: a unifying triple network model," Trends in cognitive sciences, vol. 15, no. 10, pp. 483 -- 506, 2011. [38] Anvar Kurmukov, Marina Ananyeva, Yulia Dodonova, Boris Gutman, Joshua Faskowitz, Neda Jahanshad, Paul Thompson, and Leonid Zhukov, "Classifying pheno- types based on the community structure of human brain networks," in Graphs in Biomedical Image Analysis, Computational Anatomy and Imaging Genetics, pp. 3 -- 11. Springer, 2017. [39] Hyekyoung Lee, Dong Soo Lee, Hyejin Kang, Boong- Nyun Kim, and Moo K Chung, "Sparse brain network IEEE Transac- recovery under compressed sensing," tions on Medical Imaging, vol. 30, no. 5, pp. 1154 -- 1165, 2011. [40] Orhan Fırat, Mete Ozay, Itır Onal, Ilke Oztekiny, and Fatos¸ T Yarman Vural, "Functional mesh learning for pattern analysis of cognitive processes," in Cognitive Informatics & Cognitive Computing (ICCI* CC), 2013 12th IEEE International Conference on. IEEE, 2013, pp. 161 -- 167. [41] Abdullah Alchihabi, Baran B Kivilicim, Sharlene D Newman, and Fatos T Yarman Vural, "A dynamic net- work representation of fmri for modeling and analyz- ing the problem solving task," in Biomedical Imaging (ISBI 2018), 2018 IEEE 15th International Symposium on. IEEE, 2018, pp. 114 -- 117. [42] Itir Onal, Mete Ozay, and Fatos T Yarman Vural, "Mod- eling voxel connectivity for brain decoding," in Pat- tern Recognition in NeuroImaging (PRNI), 2015 Inter- national Workshop on. IEEE, 2015, pp. 5 -- 8. [43] Chong-Yaw Wee, Pew-Thian Yap, Daoqiang Zhang, Lihong Wang, and Dinggang Shen, "Constrained sparse functional connectivity networks for mci classi- fication," Medical Image Computing and Computer- Assisted Intervention -- MICCAI 2012, pp. 212 -- 219, 2012. [44] Pedro A Vald´es-Sosa, Jose M S´anchez-Bornot, Agust´ın Lage-Castellanos, Mayrim Vega-Hern´andez, Jorge Bosch-Bayard, Lester Melie-Garc´ıa, and Erick Canales- Rodr´ıguez, "Estimating brain functional connectivity with sparse multivariate autoregression," Philosophical Transactions of the Royal Society of London B: Biolog- ical Sciences, vol. 360, no. 1457, pp. 969 -- 981, 2005. [45] Mikail Rubinov and Olaf Sporns, "Complex network measures of brain connectivity: uses and interpreta- tions," Neuroimage, vol. 52, no. 3, pp. 1059 -- 1069, 2010. [46] Hae-Jeong Park and Karl Friston, "Structural and func- tional brain networks: from connections to cognition," Science, vol. 342, no. 6158, pp. 1238411, 2013. [47] Jonathan D Power, Damien A Fair, Bradley L Schlaggar, and Steven E Petersen, "The development of human functional brain networks," Neuron, vol. 67, no. 5, pp. 735 -- 748, 2010. [48] Danielle Smith Bassett and ED Bullmore, "Small-world brain networks," The neuroscientist, vol. 12, no. 6, pp. 512 -- 523, 2006. [49] Mikail Rubinov and Olaf Sporns, "Weight-conserving characterization of complex functional brain networks," Neuroimage, vol. 56, no. 4, pp. 2068 -- 2079, 2011. [50] Andrew Zalesky, Alex Fornito, and Edward T Bull- more, "Network-based statistic: identifying differences in brain networks," Neuroimage, vol. 53, no. 4, pp. 1197 -- 1207, 2010. [51] Sophie Achard and Ed Bullmore, "Efficiency and cost of economical brain functional networks," PLoS com- putational biology, vol. 3, no. 2, pp. e17, 2007. [52] Manuel Blesa, Gemma Sullivan, Devasuda Anblagan, Emma J Telford, Alan J Quigley, Sarah A Sparrow, Ahmed Serag, Scott I Semple, Mark E Bastin, and James P Boardman, "Early breast milk exposure modi- fies brain connectivity in preterm infants," NeuroImage, 2018. [53] Urs Braun, Sarah F Muldoon, and Danielle S Bassett, "On human brain networks in health and disease," eLS, 2015. [54] Arman Afrasiyabi, Itir Onal, and Fatos T Yarman Vu- ral, "A sparse temporal mesh model for brain decod- ing," in Cognitive Informatics & Cognitive Computing (ICCI* CC), 2016 IEEE 15th International Conference on. IEEE, 2016, pp. 198 -- 206. [55] Francisco Pereira, Tom Mitchell, and Matthew Botvinick, "Machine learning classifiers and fmri: a tutorial overview," Neuroimage, vol. 45, no. 1, pp. S199 -- S209, 2009. [56] David D Cox and Robert L Savoy, "Functional magnetic resonance imaging (fmri)brain reading: detecting and classifying distributed patterns of fmri activity in human visual cortex," Neuroimage, vol. 19, no. 2, pp. 261 -- 270, 2003. [57] Nathalie Tzourio-Mazoyer, Brigitte Landeau, Dimitri Papathanassiou, Fabrice Crivello, Olivier Etard, Nico- las Delcroix, Bernard Mazoyer, and Marc Joliot, "Au- tomated anatomical labeling of activations in spm using a macroscopic anatomical parcellation of the mni mri single-subject brain," Neuroimage, vol. 15, no. 1, pp. 273 -- 289, 2002. [58] Sky McKinley and Megan Levine, "Cubic spline inter- polation," College of the Redwoods, vol. 45, no. 1, pp. 1049 -- 1060, 1998. [59] Matteo Frigo and Steven G Johnson, "Fftw: An adaptive software architecture for the fft," in Acoustics, Speech and Signal Processing, 1998. Proceedings of the 1998 IEEE International Conference on. IEEE, 1998, vol. 3, pp. 1381 -- 1384. [60] William T Cochran, James W Cooley, David L Favin, Howard D Helms, Reginald A Kaenel, William W Lang, George C Maling, David E Nelson, Charles M Rader, and Peter D Welch, "What is the fast fourier trans- form?," Proceedings of the IEEE, vol. 55, no. 10, pp. 1664 -- 1674, 1967. [61] Kiyotoshi Matsuoka, "Noise injection into inputs in back-propagation learning," IEEE Transactions on Sys- tems, Man, and Cybernetics, vol. 22, no. 3, pp. 436 -- 440, 1992. [62] Russell Reed, Seho Oh, and RJ Marks, "Regularization using jittered training data," in Neural Networks, 1992. IJCNN., International Joint Conference on. IEEE, 1992, vol. 3, pp. 147 -- 152. [63] Baran Baris Kivilcim, Itir Onal Ertugrul, and Fatos T Yarman Vural, "Modeling brain networks with ar- tificial neural networks," in Graphs in Biomedical Im- age Analysis and Integrating Medical Imaging and Non- Imaging Modalities, pp. 43 -- 53. Springer, 2018. [64] Itir Onal, Mete Ozay, and Fatos Tunay Yarman Vural, "A hierarchical multi-resolution mesh network for brain decoding," arXiv preprint arXiv:1607.07695, 2016. [65] Ulrik Brandes, "A faster algorithm for betweenness cen- trality," Journal of mathematical sociology, vol. 25, no. 2, pp. 163 -- 177, 2001. [66] Giorgio Fagiolo, "Clustering in complex directed net- works," Physical Review E, vol. 76, no. 2, pp. 026107, 2007. [67] Rong-En Fan, Kai-Wei Chang, Cho-Jui Hsieh, Xiang- Rui Wang, and Chih-Jen Lin, "Liblinear: A library for large linear classification," Journal of machine learning research, vol. 9, no. Aug, pp. 1871 -- 1874, 2008. [68] Alain Dagher, Adrian M Owen, Henning Boecker, and David J Brooks, "Mapping the network for planning: a correlational pet activation study with the tower of lon- don task," Brain, vol. 122, no. 10, pp. 1973 -- 1987, 1999. [69] JB Rowe, AM Owen, IS Johnsrude, and RE Passing- ham, "Imaging the mental components of a planning task," Neuropsychologia, vol. 39, no. 3, pp. 315 -- 327, 2001. [70] Mingrui Xia, Jinhui Wang, and Yong He, "Brainnet viewer: a network visualization tool for human brain connectomics," PloS one, vol. 8, no. 7, pp. e68910, 2013. [71] Louise H Phillips, "The role of memory in the tower of london task," Memory, vol. 7, no. 2, pp. 209 -- 231, 1999. [72] Fred Phillips, "A research note on accounting students' epistemological beliefs, study strategies, and unstruc- tured problem-solving performance," Issues in Account- ing Education, vol. 16, no. 1, pp. 21 -- 39, 2001. [73] KL Kafer and Michael Hunter, "On testing the face validity of planning/problem-solving tasks in a normal population," Journal of the International Neuropsycho- logical Society, vol. 3, no. 2, pp. 108 -- 119, 1997. [74] Matthew L Stanley, Sean L Simpson, Dale Dagenbach, Robert G Lyday, Jonathan H Burdette, and Paul J Lauri- enti, "Changes in brain network efficiency and working memory performance in aging," PLoS One, vol. 10, no. 4, pp. e0123950, 2015. (a) Planning Brain Network. Fig. 5: Regions with the highest network measures for Planning (Top) and Execution (Bottom). (b) Execution Brain Network. (a) Anatomical regions with higher node betweenness during planning. (b) Anatomical regions with higher node betweenness during execution. Fig. 6: Anatomical regions with higher node betweenness during planning (Top) and during execution (Bottom). (a) Anatomical regions with higher node out-strength during planning. (b) Anatomical regions with higher node out-strength during execution. Fig. 7: Anatomical regions with higher node out-strength during planning (Top) and during Execution (Bottom). (a) Anatomical regions with higher local efficiency and clustering coefficient during planning. (b) Anatomical regions with higher transitivity during planning. Fig. 8: Anatomical regions with higher local efficiency and clustering coefficient (Top) and higher transitivity (Bottom) during planning.
1509.01677
1
1509
2015-09-05T08:31:44
Combinatorics of Place Cell Coactivity and Hippocampal Maps
[ "q-bio.NC" ]
It is widely accepted that the hippocampal place cells' spiking activity produces a cognitive map of space. However, many details of this representation's physiological mechanism remain unknown. For example, it is believed that the place cells exhibiting frequent coactivity form functionally interconnected groups---place cell assemblies---that drive readout neurons in the downstream networks. However, the sheer number of coactive combinations is extremely large, which implies that only a small fraction of them actually gives rise to cell assemblies. The physiological processes responsible for selecting the winning combinations are highly complex and are usually modeled via detailed synaptic and structural plasticity mechanisms. Here we propose an alternative approach that allows modeling the cell assembly network directly, based on a small number of phenomenological selection rules. We then demonstrate that the selected population of place cell assemblies correctly encodes the topology of the environment in biologically plausible time, and may serve as a schematic model of the hippocampal network.
q-bio.NC
q-bio
Combinatorics of Place Cell Coactivity and Hippocampal Maps Andrey Babichev1,2, Daoyun Ji3, Facundo M´emoli4 and Yuri Dabaghian1,2∗ 1Jan and Dan Duncan Neurological Research Institute, Baylor College of Medicine, Houston, TX 77030, 2Department of Computational and Applied Mathematics, Rice University, Houston, TX 77005 3Department of Neuroscience, Baylor College of Medicine, Houston, TX 77030, 4Department of Mathematics, Ohio State University, Columbus, OH 43210 E-mail: [email protected] (Dated: July 16, 2021) It is widely accepted that the hippocampal place cells' spiking activity produces a cognitive map of space. However, many details of this representation's physiological mechanism remain unknown. For example, it is believed that the place cells exhibiting frequent coactivity form functionally inter- connected groups -- place cell assemblies -- that drive readout neurons in the downstream networks. However, the sheer number of coactive combinations is extremely large, which implies that only a small fraction of them actually gives rise to cell assemblies. The physiological processes responsi- ble for selecting the winning combinations are highly complex and are usually modeled via detailed synaptic and structural plasticity mechanisms. Here we propose an alternative approach that allows modeling the cell assembly network directly, based on a small number of phenomenological selection rules. We then demonstrate that the selected population of place cell assemblies correctly encodes the topology of the environment in biologically plausible time, and may serve as a schematic model of the hippocampal network. 5 1 0 2 p e S 5 ] . C N o i b - q [ 1 v 7 7 6 1 0 . 9 0 5 1 : v i X r a 2 I. INTRODUCTION The mammalian hippocampus plays a major role in spatial learning by encoding a cognitive map of space -- a key component of animals' spatial memory and spatial awareness [1, 2]. A remarkable property of the hippocampal neurons -- the place cells -- is that they become active only in discrete spatial regions -- their respective place fields [3] (Fig. 1A). A number of studies have demonstrated that place cell activity can represent the animal's current location [4, 5], its past navigational experience [6, 7], and even its future planned routes [8, 9]. Numerical simulations suggest that a population of place cells can also encode a global spatial connectivity map of the entire environment [10 -- 12]. Hence, it is believed that the large-scale hippocampal representation of space emerges from integrating the information provided by the individual place cells, although the details of this process remain poorly understood. Experimental studies point out that the hippocampal map is topological in nature, i.e., it is more similar to a subway map than to a topographical city map [13 -- 16]. In [10] we proposed a computational approach for modeling its structure based on several remarkable parallels between the notions of hippocampal physi- ology and algebraic topology. For example, Cech's theorem asserts that the topological structure of a space X can be deduced from the pattern of overlaps between regions that cover it (for details see [17] and Methods in [10]). The argument is based on building a special simplicial complex N, each n-dimensional simplex of which corresponds to a nonempty overlap of n + 1 covering regions, and demonstrating that the topological signatures of N and X are same [17]. Since the place cells' spiking activity induces a covering of the envi- ronment by the place fields, called a place field map (Fig. 1B), Cech's theorem suggests that the place cells' coactivity (Fig. 1C), which marks the overlaps of the place fields, may be used by the brain to represent the topology of the environment. In [10 -- 12] it was demonstrated that place cell coactivity can in fact be used to construct a "temporal" analogue of the nerve complex, T , the simplexes of which, σ = [c1, c2, ..., ck], cor- respond to the combinations of coactive place cells, c1, c2, ..., ck (Fig. 1D). Using the methods of persistent homology [18, 19] it was shown that the topological structure of T captures the topological properties of the environment, if the range of place cell spiking rates and place field sizes happen to parallel biological values derived from animal experiments [10, 11]. However, it remained unclear whether it is possible to physically implement this information in the (para)hippocampal network. On the one hand, electrophysiological studies suggest that place cells showing repetitive coactivity tend to form so-called cell assemblies -- functionally interconnected neuronal groups that synaptically drive a readout neuron in the downstream networks [20 -- 23] -- which may be viewed as "physiological simplexes" implementing T . On the other hand, the place cell combinations of T are much too numerous to be implemented physiologically. In a small environment, c.a. 1 × 1 m, thousands of place cells are active and the activity of 50 -- 300 of them is near maximal level at every given location [21]. The number of combinations of hundreds of coactive cells in an ensemble of thousands is unrealistically large, 3000 ∼ 10200. The number of cells in most parahippocampal regions, which may potentially comparable to C100 serve as readout neurons, is similar to the number of place cells [24]. This implies that only a small fraction of coactive place cell groups may be equipped with readout neurons, i.e., that the cell assemblies may encode only a small part of the place cell coactivities -- those which represent a certain "critical mass" of spatial connections. Physiologically, the place cell assemblies emerge from dynamically changing constellations of synaptic connections and are commonly studied in terms of the synaptic and structural plasticity mechanisms [25 -- 29]. For a better qualitative understanding of the qualitative properties of the cell assembly network and practical modeling of the hippocampus' functions, we propose a biologically plausible empirical approach that allows selecting the most prominent combinations of coactive place cells directly and demonstrate that the resulting population of cell assemblies is sufficient for representing the topology of the environment. 3 FIG. 1: Place fields and place cells. (A) The blue, green and brown dots, corresponding to the spikes produced by three different place cells, form well-defined spatial clusters, which represent their respective place fields. Spikes are positioned in space according to the animal's coordinates at the time of spiking. (B) A place field map produced by an ensemble of 300 place cells with mean peak firing rate f = 20 Hz and mean place field size s = 14 cm located in a 1 × 1 m environment. (C) A short time segment of the spike trains produced by three place cells. The periods of the cells' coactivity, marked by dashed lines, indicate overlap of their respective place fields (panel A): cells c1 and c2 are coactive in the region 12, cells c1, c2 and c3 are co-active in the region 123. (D) A simplex σ123 represents schematically the spatial connectivity encoded by the coactivity of cells c1, c2 and c3. Its 1D edges correspond to pairwise coactivity, e.g. σ12 represents the coactivity of cells c1 and c2. II. THE METHODS Mathematically, the task of identifying a subpopulation of coactive place cell combinations corresponds to selecting according to biologically motivated criteria a subcomplex T0 of the full coactivity complex T . The cell assemblies correspond to the maximal simplexes of T0, (i.e., the ones that are not subsimplexes of any other simplex), in contrast with the maximal simplexes of the coactivity complex, T , which can represent any largest combinations of coactive cells. The "cell assembly complex," T0, should satisfy several general requirements: I. Effectiveness. In the readercentric approach [21], each cell assembly drives a coincidence detector readout neuron in the downstream brain regions. Since the number of the readout neurons is comparable to the number of place cells, the total number of the maximal simplexes in T0, Nmax(T0), should be comparable to the number of its vertexes, Nc(T0), Nmax(T0) ≈ Nc(T0). However, the algorithm for selecting T0 should reduce only the number of coactive place cell combinations and not the place cells themselves, meaning that the number of vertexes in T and in T0 should not differ significantly. In mathematical literature, the number of k-dimensional simplexes of a simplicial complex is usually denoted as fk, and the list f = ( f0, f1, , fd) is referred to as the complex's f -vector [30]. However, since in neuroscience literature the letter f is often used to denote firing rates, we denote the number of k-dimensional simplexes by Nk. As a shorthand notation, we use Nmax to denote the number of the maximal simplexes and Nc the number of 0-dimensional simplexes in a given complex. II. Parsimony. To avoid redundancy, only a few cell assemblies should be active at a given location. Conversely, the rat's movements should not go unnoticed by the hippocampal network, i.e., the periods during which all place cell assemblies are inactive should be short. III. Contiguity. A transition of the spiking activity from one cell assembly σi to another σi+1 occurs when some cells in σi shut off and a new group of cells activates in σi+1 (see Suppl. Movies). The larger is the subassembly σi,i+1 = σi ∩ σi+1 that remains active during this transition (i.e., the more cells are shared by σi and σi+1) the more contiguous is the representation of the rat's moves and hence of the space in which it moves. The overlap between a pair of consecutively active simplexes can be characterized by a contiguity 4 FIG. 2: Topological loops: each horizontal bar represents the timeline of a topological cycle in T (T): 0D loops (connectivity components) and the 1D loops. Most cycles last over a short time before disappearing. A few remaining, persistent loops express stable topological information that may correspond to physical obstacles in the rat's environment. The time required for the correct number of cycles to appear is interpreted as the minimal time Tmin required for the rat to learn the environment. The environment used in these simulations (Figure 1B) is topologically connected (b0 = 1), and has one central hole (b1 = 1). Thus, the topo- logical barcode of this environment -- the list of Betti numbers (b0, b1, b2,...) -- is (1,1, 0, 0, ). The last spurious loop (blue 1D loop) disappears at about Tmin = 4.6 minutes, which is the learning time in this case. index ξ = dim(σi ∩ σi+1) √ dim(σi) dim(σi+1) , which assumes the maximal value ξ = 1 for coinciding cell assemblies and ξ = 0 for disjoint ones. In constructing a cell assembly complex, we expect that the mean contiguity over the simplexes in T0 should not be lower than in T . IV. Completeness. The cell assembly complex T0 should capture the correct topological signatures of the environment, such as obstacles, holes, and boundaries. For example, the lowest dimensional 0D and 1D loops in T0 represent, respectively, the piecewise and the path connectivity of the environment, as they are captured by the place cell coactivity. This information should emerge from the "topological noise" in a biologically plausible time period, comparable to the time required to obtain this information via the full complex, T (see [10, 11] and Fig. 2). Place cell spiking is modeled as a time-dependent Poisson process with spatially localized rate λc(r) = fce − (r−rc)2 s2 c , where r is a point in the environment, fc is the maximal firing rate of a place cell c, and sc defines the size of the corresponding place field centered at rc [31]. In a familiar environment, the place fields are stable, that is, the parameters fc, sc and rc remain constant [32, 33]. In our simulations, all computations were performed for ten place cell ensembles, each containing 300 neurons with an ensemble mean maximal firing rate of 20 Hz and a mean place field size of 30 cm. The place field centers in each ensemble were randomly scattered across the environment and most quantities reported in the Results were averaged over ten place field configurations. Spatial map. We simulated the rat's movements through a small (1× 1 m) planar environment (Fig. 11), similar to the arenas used in typical electrophysiological experiments (see Methods in [10]) over T = 25 minutes -- the duration of a typical "running session." The spatial occupancy rate of the rat's trajectory (i.e., the histogram of times spent at a particular location) and the frequency of the place cells' activity are shown on Fig. 3A,B. The mean speed of the rat is 20 cm/sec, so that turning around the central obstacle takes about 7 seconds. By analogy with the place fields, we designate the spatial domain where a combination of place cells comprising a simplex σ is active as its simplex field, sσ (Fig. 3C). If the simplex corresponds to a cell assembly, then sσ may also be referred to as the cell assembly field. Similarly to the place fields and the place field map (Fig.1B), the collection of all simplex fields forms a simplex field map and the cell assembly fields form a cell assembly map ((Fig. 3C). These maps provide a better "geometric proxy" for the rat's cognitive map because they illustrate both the activity and the coactivity of the individual place cells 5 FIG. 3: Spatial maps. (A) Occupancy of spatial locations in a 1 × 1 m environment -- a 2D histogram of the time spent by the animal in different locations. (B) Frequency of place cells' spiking: each dot marks the location of a place cell's center rc and indicates the corresponding appearance rate according to the colorbar. Higher appearance rates appear in the domain where the spatial occupancy is higher. (C) Simplex field map. The place field map for the same place cell ensemble is shown in Fig. 1B. (D) Spatial distribution of the frequency of the maximal simplexes' appearances. Notice that, since place cells with higher appearance rates tend to produce higher order cell assemblies, which, in turn, have lower appearance rates, the spatial distribution of rates on B. and D. are complementary. ((Fig. 3C-D). In the following, the structure of these maps will be used to discuss our selection algorithms. If the distinction between a cell assembly map and a simplex map is not essential, it will be referred to as a space map. Population activity. To define the population code [34] of place cell combinations, we construct place cell activity vectors by binning spike trains into 1/4 sec long time bins (for a physiological justification of this value see [11, 35]). If the time interval T splits into n such bins, then the activity vector of a cell c is mc(T) = [mc,1, , mc,n], where mc,k specifies how many spikes were fired by c in the kth time bin. The components of mc, normalized by the total number of spikes, Mc, define spiking probabilities, pc,k = mk/Mc [36]. A stack of activity vectors forms an activity raster illustrated on Fig. 4. Two cells, c1 and c2, are coactive over a certain time period T, if the dot product of their activity vectors does not vanish, mc1(T) · mc2(T) (cid:44) 0. The component-wise or Hadamard product of two activity vectors mc1,c2 = mc1 (cid:12) mc2 = [mc1,1mc2,1, mc1,2mc2,2, , mc1,nmc2,n] defines the coactivity vector of cells c1 and c2, which can also be viewed as the activity vector of the corresponding 1D simplex σ12 = [c1, c2], mσ12 ≡ mc1,c2. Similarly, the Hadamard product of k vectors, mσ12...k = mc1,c2,...,ck = mc1 (cid:12) mc2 (cid:12) (cid:12) mck , defines the activity vector of the simplex σ12...k = [c1, c2, , ck]. For each activity vector, mσ, we also define its bit array mapping into a binary appearance vector, aσ, which indicates during which time-bins the corresponding simplex σ has made its appearance, i.e., aσ,i = 1 iff mσ,i > 0. The appearance rate, fσ(T), of a simplex σ over a time interval T, is defined as the L1 norm of its appearance vector, averaged over that time interval, fσ(T) = (1/T)Σiaσ,i. These appearance vectors and appearance rates allow distinguishing the intrinsic physiological characteris- tics of place cells' spiking, e.g., their maximal firing rate, from the frequency with which these cells activate due to the rat's movements through their respective place fields. While the maximal firing rate of a typical place cell is about 15 Hz [1], the frequency of their activation is much lower. 6 FIG. 4: An activity raster of a population of 20 place cells over 250 time bins. Each row defines the activity vector of the corresponding place cell. The color of the ticks indicates the number of spikes con- tained in the corresponding bin, according to the col- orbar on the right. At every time step, the nonempty bins in the vertical column define the list of cur- rently active cells, i.e., the active simplex σt. During the time interval T, marked by the two vertical blue lines, cell c16 is coactive with c18 but not coactive with c14. III. RESULTS The simulated ensembles of 300 place cells in the environment shown on Fig. 1B produced a coactivity complex T with about Nmax = 1000 maximal simplexes. Despite the high dimensionality of these simplexes (up to D = 35, mean ¯D = 17), the characteristic dimensionality of a facet shared by two consecutively active simplexes, σi and σi+1, is relatively low, so that the mean contiguity of T is ξ = 0.6. This implies that, geometrically, if the simplexes of T are viewed as multidimensional tetrahedrons, the selected complex, T0(θ), assumes a highly irregular shape (Fig. S1A). More importantly, nearly 100% of the maximal simplexes appeared only once during the entire 25 minute period of navigation, i.e., a typical maximal simplex's appearance rate is low, fσ ∼ 10−3 Hz. How- ever, a typical vertex activated about 200 times or every seven seconds, suggesting that some of the lower dimensional subsimplexes may be better candidates for forming cell assemblies. Is it then possible to build a cell assembly complex T0 by discarding the high-dimensional maximal simplexes with low appearance rates and retaining their subsimplexes that appear more frequently? We tested this hypothesis by identifying the combinations σ whose coactivity exceeds a certain threshold fσ > θ, and studied the properties of the resulting simplicial complex as a function of θ (Fig. 5A). First we observed that, as soon as the appearance threshold is introduced (θ (cid:38) 10−3 Hz), the high dimen- sional simplexes start braking up, releasing large numbers of lower dimensional subsimplexes: the number of k-dimensional subsimplexes in a n-dimensional simplex grows as combinatorial coefficient Ck+1 n+1, e.g., 18 ≈ 44, 000. As a result, the complex T0(θ) rapidly inflates. As θ increases further for n = 17 and k = 7, C8 (θ > 0.04), the number of "passing" simplexes decreases, and T0(θ) begins to shrink in all dimensions (i.e., ND(θ1) > ND(θ2) for θ1 < θ2, for all D, (Fig. 5B). Despite this, the number of maximal simplexes remains high: Nmax = 30× Nc at θ = 0.04 Hz, Nmax = 7× Nc at θ = 0.07 Hz and Nmax = 3× Nc for the highest tested threshold, θ = 0.1 Hz (Fig. 5B), while their characteristic dimensionality drops from D = 17 to D = 7 at θ = 0.04 Hz and to D = 3 at θ = 0.1 Hz. The mean contiguity index for this range of thresholds remains close to ξ = 0.7, indicating that the degree of overlap between the selected combinations of place cells is higher than in the original coactivity complex. However, raising the passing threshold θ quickly destroys the geometric integrity of the resulting com- plex's spatial map. As shown on Fig. S2, for θ = 0.05 Hz, only ∼ 50% of the environment is covered by the remaining simplex fields, and for θ = 0.07 Hz the simplex map barely retains its one-piece connectedness: in some cases the complex T0 splits in two (the corresponding Betti numbers, b0, are listed in Table S1, for an illustration see Fig. S3). For θ = 0.1 Hz, the complex fragments into multiple components (mean b0 ∼ 7) that are riddled with holes: the Betti numbers bn>0 indicate the presence of hundreds of stable loops in higher dimensions. Thus, even if the coactive place cell combinations selected at θ ≥ 0.05 Hz could be supplied with readout neurons and would form cell assemblies, the resulting cell assembly network would not encode the correct spatial connectivity. An additional problem is that reducing the order of the assemblies violates the "assembly code" for spatial locations: every time several subsimplexes σi are selected from a high-order maximal simplex σ, 7 FIG. 5: A direct selection of the simplexes by appearance rates. (A) In the original coactivity complex T (θ = 0), the maximal simplexes σmax appear on average but once during the entire observation period, resulting in low appearance rates ( fσ < 10−3 Hz, blue line). Imposing four different thresholds θ (color coded) raises the appearance rates of the selected maximal simplexes almost uniformly in all dimensionalities. (B) Cumulative distribution of the number of maximal simplexes Nmax over the selected simplexes' dimension. In the T (θ = 0) case Nmax exceeds the number of vertexes (Nc = 300, black horizontal line) by almost an order of magnitude. Small threshold values result in an explosive increase of Nmax which then begins to decrease for θ > 0.04 Hz, remaining significantly higher than Nc for all four tested values of θ. (C) The histograms of the maximal simplexes' dimensionalities fit with normal distribution. The high mean dimensionality ( ¯D = 17) observed in the T (θ = 0) case reduces to ¯D = 2.2 for θ = 0.1 Hz. The width of the distributions is about 50% of ¯D. (D) The histograms of the number of the coactive maximal simplexes, fit to an exponential distribution demonstrate that the typical number of coactive simplexes is large, β > 10. All values are averaged over ten place field maps generated by ten place cell ensembles with the same mean peak firing rate and mean place field size. several overlapping simplex fields sσi are produced in place of a single sσ. As a result, the parsimony of the representation is compromised: a location that was previously represented by a single simplex becomes represented by a few of its subsimplexes (Fig. S4A-B). Fig. 5D shows a histogram of the numbers of simultaneously active maximal simplexes in T0: although most of the time only a few maximal simplexes are active, a coactivity of many of them (n > 25) is not uncommon. Conversely, while most of the time -- on average 84% for the selected place cell ensembles -- at least one simplex is active, longer inactivity periods are observed as described by double exponential distributed with the rate β ≈ 3.5 sec (Fig. S2). Overall, since most of the T0-requirements listed in the Methods fail, we are led to conclude that the most straightforward selection rule, based on selecting high appearance rates, does not produce the de- sired tradeoff between the order of the assemblies, the frequency of their appearances, and the quality of topological representation of the environment. This failure motivates the search for alternative methods. Method I. To produce a more detailed approach to selecting coactive cell combinations, we observe that place fields are typically convex planar regions, and hence the existence of higher order overlaps between them actually follows from the lower order overlaps. According to Helly's theorem, a collection of n > D + 1 convex D-dimensional regions in Euclidean space RD will necessarily have a nonempty common intersection, if the intersection of every set of D + 1 regions is nonempty (see [37, 38] and Fig. S4D). From the perspective of Cech theory, this implies that if n convex regions which cover a D-dimensional space contribute all the combinatorially possible D-dimensional simplexes to the nerve complex, then they also provide all the higher (up to n−1) dimensional simplexes to it. In a planar (D = 2) environment, this implies that a set of four or more place fields has a common intersection, if any three of them overlap. Moreover, although mathematically it is possible that three place fields exhibit pairwise, but not triple overlap, the probability of such an occurrence is low (Fig. S4C). A direct computational verification shows that if a triple of place cells demonstrates pairwise coactivity, then, in over 90% of cases, it also correctly encodes a triple spatial overlap. In other words, a "clique" of pairwise coactivities indicates the overlaps of all higher orders, which implies that the spatial connectivity graph GN whose vertexes correspond to the place fields 8 FIG. 6: Selecting maximal simplexes via the pairwise coactivity threshold (Method I). (A) The appearance rates of the maximal simplexes computed for four different pairwise appearance rate thresholds θ decrease as a function of their dimensionality. The values at D = 1 correspond to the value of the threshold imposed on the links' appearance rate. (B) Cumulative distribution of the numbers of maximal simplexes, Nmax, over the selected simplexes' dimension. The numbers of cells Nc for each threshold value are shown by horizontal lines. The tendency of the maximal simplexes to outnumber the vertexes Nmax > Nc, characteristic for small values of θ, is reversed around θ = 0.07 Hz, where Nmax and Nc level out. (C) The histograms of the maximal simplexes' dimensionalities fit with normal distribution. The mean dimensionalities are similar to the ones produced by the previous selection method. The width of the distributions is about 50% of ¯D. (D) The histogram of the number of coactive maximal simplexes, fit to an exponential distribution, shows that the expected number of coactive simplexes (β ∼ 4) is significantly lower than in the previous selection method. The procedure of averaging over the place field maps is the same. and links represent pairwise overlaps, encodes most simplexes in the nerve complex N. As a reminder, a clique of an undirected graph is a set of pairwise connected vertices. From the combi- natorial perspective, a clique and a simplex have the same defining property: any subset of a simplex is its subsimplex and any subset of a clique is its subclique; a maximal clique is the one that is contained in no other clique. Hence, each graph defines its own "clique complex," the k-dimensional simplexes of which corresponds to the graph's cliques with k + 1 vertices [39]. The observation that the nerve complex induced from the place field map can be approximated by the clique complex of the place field pairwise connectivity graph, suggests that the corresponding coactivity complex T can also be built based only on pairwise, rather than higher-order, coactivities. This approach is well justified physiologically, since pairwise coactivity detector pairs of synapses are commonly observed [40, 41]. The rule for defining the temporal analogue of GN -- the relational graph GT -- is straightforward: a pair of vertexes is connected in GT if the corresponding cells ci and c j are coactive. Thresholding pairwise coactivity rates according to the rule if fci,c j ≥ θ if fci,c j < θ. (1) 1 0 Ci j = allows constructing a family of relational graphs GT (θ) over the pairs of place cells with high coactivity. The higher the threshold is, the sparser its connectivity matrix Ci j and the smaller the number of maximal cliques and hence of maximal simplexes in the corresponding clique complex. Since in the following the graph GN will not be used we will suppress the subscript "T " in the notation for GT . We studied the relational graphs G(θ) and their respective clique complexes T0(G(θ)) ≡ T0(θ) as a func- tion of θ. First, we observed that the appearance rates of the maximal simplexes in T0(θ) become sensitive to the simplexes' dimensionality (Fig. 6A), implying that this selection procedure in effect attributes differ- ent thresholds to simplexes of different dimensions by using only one free parameter θ. Second, the size of T0(θ) is not as large as before. As shown on Fig. 6B, even for a relatively low threshold θ = 0.05 Hz, the number of maximal simplexes exceeds the number of cells only marginally. For higher thresholds, this number steadily decreases: ND(θ1) < ND(θ2) for θ1 > θ2 > 0.04 and D > 3, though in lower dimensions 9 FIG. 7: Figure 7. Correction algorithms. (A) A spatial projection of the 2D skeleton of T (θ) shows gaps and holes that compromise, respectively, the piecewise and path connectivity of T (θ). If the links across the gaps and holes of T (θ) are restored, then its correct connectivity structure may be regained. (B) A 'hole' produced by five connected vertexes is closed by restoring some of the previously discarded crosslinks. (C) A projection of the relational graph G into the environment, shown in grey. The edges added across the gaps are shown in red and the edges added to fill the holes are shown in green. The vertexes that are left disconnected due to low appearance rates of the edges connecting them to other vertexes are shown by blue dots. (D) A spatial map of the resulting "patched" 2D skeleton of T (θ). The parameter values are ng = 15, mh = 10, and the lowered threshold for reintroducing the missing links is θ = 50. (1 ≤ D ≤ 3) this number may increase. The characteristic contiguity ranges between ξ = 0.65 at θ = 0.05 Hz to ξ = 0.72 at θ = 0.14 Hz, which is higher than the value produced by the direct simplex selection method. Geometrically, this implies that the collection of maximal simplexes selected by pairwise thresh- old selection is more aggregated than the collection produced via direct simplex selection, i.e., the resulting complex T0(θ) is geometrically more similar to a "simplicial quasimanifold" (see Fig. S1B). However, the number of place cells Nc drops as a result of discarding too many links with low appearance rate: Nc = 290 at θ = 0.05 Hz and Nc = 100 at θ = 0.14 Hz. At θ = 0.1 Hz number of cells levels out with the number of maximal simplexes, Nmax ∼ Nc = 260. As before, raising the coactivity threshold degrades the spatial map. At θ > 0.07 Hz the simplex fields no longer cover the environment and at θ > 0.1 Hz the map fragments into pieces (Fig. S5). However the resulting complex exhibits a much more regular topological behavior: the correct signature (b0 = 1, b1 = 1, b2 = 0, b3 = 0, ...) in T0(θ) appears at θ = 0.05 Hz. The higher order Betti numbers (bn≥2) remain trivial at still higher θs (Table S2A), even though the connectedness and path connectivity of the environment (b1 and b0) become misrepresented. This improvement of the behavior of T0(θ) suggests that, despite all the shortcomings, the link-selection strategy may lead to a successful model of the place cell assembly network. After all, it is not surprising that a single selection rule does not resolve all the aspects of the cell assembly formation. Yet if it captures the essence of the process, it should be possible to correct or to adjust its outcome. For example, one of the difficulties faced by the coactivity selection algorithm is that, for high θ, T0(θ) may brake into several pieces. However, the gaps between them are small. Thus, if a few discarded edges of the relational graph that originally bridged these gaps are retained, then the connectedness of T0(θ) may be spared (Fig. 7A). Similarly, a "hole" in the relational graph is a linear chain of edges, connected tail to tail, with no shortcuts. However, if the links with the lower appearance rate ( f ≥ θh, θh < θ) that span across the hole exist at θ = 0, then they also can be restored (Fig. 7B). This may remove the non-contractible chains of 1D simplexes in T0(θ) that compromised its path connectivity (Fig. 7C,D). Thus, we implemented the following two rectification algorithms: 1. Filling gaps: find pairs of vertexes va and vb separated in G(θ) by more than ng edges and then test whether these vertexes are connected directly by links (from G(θ = 0)) whose appearance rate exceeds a lower threshold θg < θ. If such links exist, add them to G(θ) (red lines on Fig. 7C). 2. Closing holes: A closed chain containing mh ≥ 4 edges in G(θ), with no shortcuts, is likely to 10 FIG. 8: Statistics of the vertex degrees in relaitonal graphs. (A) The histogram of the vertex degrees k in the neighbor-controlled relational graph G(n0), computed for four different n0 (Method II) and fitted to a power law distribution P(k) ∼ k−γ. The graph demonstrates that G(n0) is a scale-free network. (B) The same distribution on the log-log scale and an independent linear fit of the powers γ. The confidence intervals of the two fits, ranging between ±0.15 and ±0.3, overlap for each case. (C) In the pairwise coactivity threshold (Method I), the histogram of the relational graph's vertex degrees is fit by negative binomial distribution, suggesting that G(θ) is similar to a random network. produce a hole in T0(θ). We identified such chains and restored the discarded cross-links whose appearance rate exceeds a lower threshold θh < θ. Thus, both rectification algorithms depend on two parameters: the length of the involved chains (ng for gaps and mh for holes) and the value of the reduced threshold θg and θh. In our numerical experiments, we found that the optimal value for the thresholds is θh = θg = 50, and the parameters range between 5 and 10 (mh) and 10 and 15 (ng). Typically, each rectification procedure is applied once or twice before the right signature of T0(θ) is achieved, and this without producing significant changes of the complex's structure, such as altering the appearance of its simplexes or increasing its size Nmax. As illustrated in the Table S2B, the correct signature in the "repaired" complex is achieved for all cases at θ = 0.07 Hz. In particular, at θ = 0.07 Hz we obtain a simplicial complex T0 with the correct signature, having Nc = 260 vertexes and about the same number of maximal simplexes, Nmax ≈ Nc. These maximal simplexes appear on average at a rate of fσ ≥ 0.07 Hz, at least during every other run of the rat around the environment, and have dimensionality D = 6. As a result, the requirements to T0 are met and the maximal simplexes of T0 may represent hippocampal place cell assemblies that together encode a map of the environment, and hence T0 itself can be viewed as the "cell assembly complex." Method II. A common feature of the appearance-rate-based selection rules is that the resulting simplicial complex reflects biases of its spatial occupancy: higher dimensional maximal simplexes concentrate over the parts of the environment where the rat appears more frequently. In particular, the relational graph shows a higher concentration of edges over the eastern segment of the environment (Fig. 7C) where the occupancy rate is highest (Fig. 3A). On the one hand, this is natural since the frequency of the place cells' spiking activity certainly does depend on the frequency of the rat's visits to their respective place fields, which therefore affects the hippocampal network's architecture [28, 29]. In fact, this argument is at the core of the classical "hippocampus as a cognitive graph" model [42, 43], which proposes that the architecture of the hippocampal network is an epiphenomenon of the place cell coactivity. On the other hand, the physiological processes that produce synaptic connections may be more autonomous. For example, the CA3 region of the hippocampus is anatomically a recurrent network of place cells whose spiking activity and synaptic architecture are dominated by the network's attractor dynamics [44 -- 46]. These considerations lead us to test an alternative method of constructing the relational graph based on selecting, for every cell, its n0 closest neighbors as defined by the pairwise coactivity rate fci,c j. Note that the resulting number of connections may be different for different cells: a cell c1 may be among the n0 closest neighbors of a cell c2, and hence c1 and c2 become connected, but the set of n0 closest neighbors of 11 FIG. 9: Figure 9. Selecting maximal simplexes via best neighbor selection (Method II). (A) The appearance rates of the maximal simplexes in the simplicial complex T (n0), computed for four different values of n0 (color coded), decrease as a function of their dimensionality. (B). Cumulative distribution of the number of maximal simplexes Nmax over the selected simplexes' dimension. For the tested values of n0, the fixed number of vertexes Nc = 300, indicated by the horizontal black line, is close to the number of maximal simplexes. For n0 = 7, the values Nmax and Nc come closest. (C). The histograms of the maximal simplexes' dimensionalities, fit to the normal distribution, indicate that for the relational graph with a similar number of links, the mean dimensionalities of the maximal simplexes are smaller than in in the complex built via the threshold-selection method. The width of the distributions is about 40% of ¯D. (D) The histogram of the number of coactive maximal simplexes, fit to an exponential distribution. An expected number of coactive simplexes ranges between β = 2 and β = 6. The procedure of averaging over the place field maps is the same. a cell c1 may not include c2, which bears a certain resemblance to the preferential attachment models [47]. As a result, the vertex degrees k of the (undirected) relational graph may differ from one another and from n0. A direct computational verification shows that k is distributed according to a power law, P(k) ∼ k−γ, where γ ranges, for different n0, between γ ∼ 2 and γ ∼ 4 (Fig. 8), which implies that G(n0) demonstrates scale-free properties [47, 48] characteristic of the hippocampal network [49, 50]. In contrast, the histogram of the vertex degrees in the threshold-controlled relational graph G(θ) may be fit with the negative binomial distribution (Fig. 8B), which indicates that G(θ) is similar to a random graph. This neighbor-selection method for building the relational graph G(n0) has a number of other immediate advantages over the threshold-controlled construction of G(θ). For example, no cells are excluded from T0 due to the low appearance of the edges connecting to them. As a result, the simplex fields are distributed more uniformly (Fig. S6), which helps capture the correct piecewise connectedness of the environment. By studying the properties of the clique complexes produced by the relational graphs G(n0) for n0 = 2, 4, 7 and 12 -- parameters chosen to produce similar numbers of edges as in the previous method -- we found that the number of maximal cliques in T0(G(n0)) is typically lower than in T0(G(θ)). The appearance rates of maximal cliques in G(n0) are more scattered and less sensitive to dimensionality than in G(θ) (Fig. 8A and Fig. S6). The number of maximal simplexes in T0(n0) remains close to the number of cells (Fig. 8B) and their dimensionality is lower than in the threshold-based selection approach (Fig. 8C). The contiguity index in all complexes ranges between to ξ = 0.67 and ξ = 0.71. The coverage of the space with the simplex fields improves with growing n0 (see Fig. S6) -- for n0 > 2 the complex T0(n0) is connected, while the behavior of b0 is more regular (see Table S3). However, the path connectivity of the complex T0(n0) remains deficient for all n0 because the number of stable spurious 1D loops remains high (Table S3). After filling the gaps and closing the holes, most complexes constructed for n0 ≥ 7 acquire correct topological signatures (Table S3), and the requirements to T0 are satisfied. Thus, the simplicial complex obtained by the neighbor selection method for n0 ≥ 7 can also be viewed as a "cell assembly complex," meaning it can serve as a formal model of the place cell assembly network with mean contiguity ξ = 0.7. IV. DISCUSSION 12 The proposed approach allows creating schematic models of the place cell assembly network -- the cell assembly complex T0 -- by controlling the basic phenomenological parameters of place cell coactivity and then relating the network's architecture to the net topological information it encodes. The coactivity com- plex T was previously used for representing the pool of place cell coactivities [10 -- 12]. Specifically, the low order (pair and triple) coactivity events were used to construct the 2D skeleton of T , and then its 0D and 1D topological loops were matched with the corresponding topological signatures of the environment. In the current study, the entire pool of place cell coactivities is used to model the full place cell assembly network, including the higher order assemblies representing both the low-dimensional spatial environment as well as high-dimensional memory space [21, 51]. The learning times Tmin estimated from the dynamics of the 0D and 1D loops in T0 remain close to the learning times computed for the full coactivity complex T (see Table S4). This implies that the selected, "core" pool of coactive place cell combinations captures the topological structure of the environment as fast and as reliably as the entire set of the place cell coactivities. We view the proposed algorithms as basic models of a more general "phenomenological" approach, one which can be further developed along several broad lines. First, the structure of the relational graph is currently deduced from the activity vectors defined over the entire navigation period T = 25 minutes. A biologically more plausible selection algorithm should be adaptive: the structure of the relational graph at a given moment of time t < T should be based only on the spiking information produced before t. Hence, in a more advanced model, the structure of the relational graph should develop in time, and in general the cell assemblies comprising T0 should be derived using synaptic and structural plasticity mechanisms. Second, the selection criteria in Methods I and II above may be individualized: the appearance threshold used to construct the relational graph can be assembly-specific, i.e. θ = θ(σ), so that the properties of the resulting network would be described in terms of the probability distribution of the threshold values across the cell assembly population. Similarly, the number of closest neighbors can be made cell-specific, n0 = n0(ci), which should permit better control over the topological properties both of the network and of the cell assembly complex. Third, threshold control can be implemented using different coactivity metrics, for instance via the pairwise correlation coefficient ρ(c1, c2) = (m1 · m2)/m1m2, (2) which would connect cells with correlated spiking (irrespective of their firing rates), in contrast with the metric (2), which does the opposite. In general, two metrics ρ and ρ(cid:48), produce relational graphs with differ- ent topologies. Nevertheless, they may produce similar or identical large-scale effects, such as generating topologically identical cell assembly complexes T0, or exhibit similar learning times, Tmin. Identifying classes of metrics that produce topologically similar results will be examined in future research. V. ACKNOWLEDGMENTS The work was supported in part by Houston Bioinformatics Endowment Fund, the W. M. Keck Founda- tion grant for pioneering research and by the NSF 1422438 grant. VI. REFERENCES 13 Ann. Rev Neurosci 24: 459-486. Press. xiv, 570 pp. p. 95: 2717-2719. [1] Best PJ, White AM, Minai A (2001) Spatial processing in the brain: the activity of hippocampal place cells. [2] O'Keefe J, Nadel L (1978) The hippocampus as a cognitive map. New York: Clarendon Press; Oxford University [3] Best PJ, White AM (1998) Hippocampal cellular activity: a brief history of space. Proc. Natl Acad. Sci. USA [4] Zhang K, Ginzburg I, McNaughton BL, Sejnowski TJ (1998) Interpreting neuronal population activity by re- construction: unified framework with application to hippocampal place cells. J Neurophysiol. 79: 1017-1044. [5] Brown EN, Frank LM, Tang D, Quirk MC, Wilson MA (1998) A statistical paradigm for neural spike train decoding applied to position prediction from ensemble firing patterns of rat hippocampal place cells. J Neurosci 18: 7411-7425. [6] Carr MF, Jadhav SP, Frank LM (2011) Hippocampal replay in the awake state: a potential substrate for memory consolidation and retrieval. Nat Neurosci 14: 147-153. [7] Derdikman D, Moser MB (2010) A dual role for hippocampal replay. Neuron 65: 582-584. [8] Pfeiffer BE, Foster DJ (2013) Hippocampal place-cell sequences depict future paths to remembered goals. Na- [9] Dragoi G, Tonegawa S (2011) Preplay of future place cell sequences by hippocampal cellular assemblies. Nature ture advance online publication. 469: 397-401. [10] Dabaghian Y, Mmoli F, Frank L, Carlsson G (2012) A Topological Paradigm for Hippocampal Spatial Map Formation Using Persistent Homology. PLoS Comput. Biol. 8: e1002581. [11] Arai M, Brandt V, Dabaghian Y (2014) The Effects of Theta Precession on Spatial Learning and Simplicial Com- plex Dynamics in a Topological Model of the Hippocampal Spatial Map. PLoS Comput. Biol. 10: e1003651. [12] Curto C, Itskov V (2008) Cell groups reveal structure of stimulus space. PLoS Comput. Biol. 4: e1000205. [13] Gothard KM, Skaggs WE, McNaughton BL (1996) Dynamics of mismatch correction in the hippocampal en- semble code for space: interaction between path integration and environmental cues. J Neurosci 16: 8027-8040. [14] Leutgeb JK, Leutgeb S, Treves A, Meyer R, Barnes CA, et al. (2005) Progressive transformation of hippocampal neuronal representations in "morphed" environments. Neuron 48: 345-358. [15] Alvernhe A, Sargolini F, Poucet B (2012) Rats build and update topological representations through exploration. Anim Cogn 15: 359-368. 10.7554/eLife.03476. [16] Dabaghian Y, Brandt VL, Frank LM (2014) Reconceiving the hippocampal map as a topological template; eLife [17] Hatcher A (2002) Algebraic topology. Cambridge; New York: Cambridge University Press. [18] Ghrist R (2008) Barcodes: The persistent topology of data. Bulletin of the American Mathematical Society 45: 61-75. 243 p. p. 719 p. p. 429-441. [19] Zomorodian AJ (2005) Topology for computing. Cambridge, UK ; New York: Cambridge University Press. xiii, [20] Harris KD, Csicsvari J, Hirase H, Dragoi G, Buzsaki G (2003) Organization of cell assemblies in the hippocam- pus. Nature 424: 552-556. [21] Buzsaki G (2010) Neural syntax: cell assemblies, synapsembles, and readers. Neuron 68: 362-385. [22] Huyck C, Passmore P (2013) A review of cell assemblies. Biol. Cybern. 107: 263-288. [23] Harris KD (2005) Neural signatures of cell assembly organization. Nat Rev Neurosci 6: 399-407. [24] Shepherd GM (2004) The synaptic organization of the brain. Oxford ; New York: Oxford University Press. xiv, [25] Wennekers T, Palm G (2009) Syntactic sequencing in Hebbian cell assemblies. Cognitive Neurodynamics 3: [26] Ghalib H, Huyck C. A Cell Assembly Model of Sequential Memory; 2007 2007. pp. 625-630. [27] Itskov V, Curto C, Pastalkova E, Buzsaki G (2011) Cell assembly sequences arising from spike threshold adap- tation keep track of time in the hippocampus. J Neurosci 31: 2828-2834. [28] Caroni P, Donato F, Muller D (2012) Structural plasticity upon learning: regulation and functions. Nat Rev 14 Neurosci 13: 478-490. [29] Chklovskii DB, Mel BW, Svoboda K (2004) Cortical rewiring and information storage. Nature 431: 782-788. [30] Gromov ML (1968) On the number of simplexes of subdivisions of finite complexes. Mathematical notes of the [31] Barbieri R, Frank LM, Nguyen DP, Quirk MC, Solo V, et al. (2004) Dynamic analyses of information encoding Academy of Sciences of the USSR 3: 326-332. in neural ensembles. Neural Comput. 16: 277-307. [32] Wilson MA, McNaughton BL (1993) Dynamics of the hippocampal ensemble code for space. Science 261: [33] Brown EN, Nguyen DP, Frank LM, Wilson MA, Solo V (2001) An analysis of neural receptive field plasticity by point process adaptive filtering. Proc. Natl Acad. Sci. USA 98: 12261-12266. [34] Pouget A, Dayan P, Zemel R (2000) Information processing with population codes. Nat Rev Neurosci 1: 125- 1055-1058. 132. 49-86. [35] Mizuseki K, Sirota A, Pastalkova E, Buzsaki G (2009) Theta oscillations provide temporal windows for local circuit computation in the entorhinal-hippocampal loop. Neuron 64: 267-280. [36] Perkel DH, Gerstein GL, Moore GP (1967) Neuronal Spike Trains and Stochastic Point Processes: I. The Single Spike Train. Biophys. J 7: 391-418. [37] Avis D, Houle ME (1995) Computational aspects of Helly's theorem and its relatives. International Journal of Computational Geometry and Applications 05: 357-367. [38] Eckhoff J (1993) CHAPTER 2.1, Helly, Radon, and Carathodory Type Theorems. In: Wills PMGM, editor. Handbook of Convex Geometry. Amsterdam: North-Holland. pp. 389-448. [39] Bandelt HJ, Chepoi V (2008) Metric graph theory and geometry: a survey. Contemporary Mathematics 453: [40] Katz Y, Kath WL, Spruston N, Hasselmo ME (2007) Coincidence detection of place and temporal context in a network model of spiking hippocampal neurons. PLoS Comput. Biol. 3: e234. [41] Brette R (2012) Computing with Neural Synchrony. PLoS Comput. Biol. 8: e1002561. [42] Muller RU, Stead M, Pach J (1996) The hippocampus as a cognitive graph. J Gen Physiol. 107: 663-694. [43] Burgess N, O'Keefe J (1996) Cognitive graphs, resistive grids, and the hippocampal representation of space. J Gen Physiol. 107: 659-662. [44] Colgin LL, Leutgeb S, Jezek K, Leutgeb JK, Moser EI, et al. (2010) Attractor-map versus autoassociation based attractor dynamics in the hippocampal network. J Neurophys. 104: 35-50. [45] Wills TJ, Lever C, Cacucci F, Burgess N, O'Keefe J (2005) Attractor dynamics in the hippocampal representation of the local environment. Science 308: 873-876. [46] Tsodyks M (2005) Attractor neural networks and spatial maps in hippocampus. Neuron 48: 168-169. [47] Barabsi A-L, Albert R (1999) Emergence of Scaling in Random Networks. Science 286: 509-512. [48] Albert R, Barabsi A-L (2002) Statistical mechanics of complex networks. Reviews of Modern Physics 74: 47-97. [49] Li X, Ouyang G, Usami A, Ikegaya Y, Sik A (2010) Scale-Free Topology of the CA3 Hippocampal Network: A Novel Method to Analyze Functional Neuronal Assemblies. Biophys. J 98: 1733-1741. [50] Bonifazi P, Goldin M, Picardo MA, Jorquera I, Cattani A, et al. (2009) GABAergic Hub Neurons Orchestrate Synchrony in Developing Hippocampal Networks. Science 326: 1419-1424. [51] Eichenbaum H, Dudchenko P, Wood E, Shapiro M, Tanila H (1999) The hippocampus, memory, and place cells: is it spatial memory or a memory space? Neuron 23: 209-226. [52] Singh G, Memoli F, Ishkhanov T, Sapiro G, Carlsson G, et al. (2008) Topological analysis of population activity in visual cortex. J Vis 8: 11 11-18. VII. SUPPLEMENTARY FIGURES 15 FIG. S1: Simplicial complexes. (A) A schematic representation of an irregular simplicial complex K, in which the number of maximal simplexes is larger than the number of vertexes (black dots). The maximal 1D simplexes are shown as blue segments, the 2D simplexes as gray triangles and the 3D simplexes as pink tetrahedrons. (B) A simplicial "quasi-manifold," Q, which has a similar number of vertexes and maximal simplexes of different dimen- sionalities. If each maximal simplex, e.g. (1, 2, 3) or (10, 7, 8, 9), corresponds to an assembly of place cells driving a readout neuron (red dots), then Q is a cell assembly complex. Dotted lines represent synaptic connections from the place cells to the readout neuron. 16 FIG. S2: The simplicial complexes T0(θ) constructed by direct selection of coactive combinations, for four different values of θ. (A) The appearance rates of simplexes, arranged from left to right according to their dimension. Each dot corresponds to a maximal simplex whose dimension is color-coded according to the colorbar on the right. (B) Spatial distribution of the dimensionalities of the selected simplexes. (C) Spatial distribution of the appearance rates of the selected simplexes. (D) The histograms of the lapse times, fit to double exponential distribution (blue line), and the value of the fitted distribution's rate β. (E) Spatial projections of the 2D skeletons of the T0(θ). Data for all panels is computed for a specific place field map for illustrative purposes. 17 FIG. S3: Low-dimensional topological features of the cell assembly complexes captured by place field map. (A) A place filed map corresponding to a singly-connected (b0 = 1, marked by the dot) complex with no non- contractible loops (b1 = 0). (B) A place field map corresponding to a complex with correct list of Betti numbers (correct topological barcode [18]): the physical hole (see Figures 1 and 3) produces one non-contractible loop (red circle, b1 = 1). (C) A map containing a spurious hole in the upper-right corner produces an extra persistent loop (the additional small circle, net b1 = 2). (D) A map containing two disconnected pieces marked by the black dots (net b0 = 2) and having no non-contractible loops (b1 = 0). (E) A map containing two pieces and one persistent spurious 1D loop. (F) The green and the brown place fields at the top connect, yielding another persistent spurious 1D loop. Compare these illustrations and the topological barcodes to the illustrations and Suppl. Movies provided in [52] 18 FIG. S4: Schematic illustrations of spatial maps. (A) Schematic representations of 3d order simplex fields, a, b, c and d, encoded by third order maximal simplexes, σa = (v1, v2, v6), σc = (v2, v3, v5), etc. (B) If the 2D simplexes are discarded and their 1D faces are retained, then the second-order simplex fields are produced, shown here as overlapping shaded regions. The original simplex fields a, b, c and d are now represented by the coactivity of three pairs, e.g., a is represented by σa,1 = (v1, v2), σa,2 = (v2, v6) and σa,3 = (v1, v6). (C). The three place fields on the left exhibit pairwise, but not triple overlap. In the generic spatial configuration shown on the right, pairwise overlapping place fields also produce a triple overlap. (D) Four pairwise overlapping convex regions in 2D produce all the higher order (triple and quadruple) overlaps. 19 FIG. S5: Simplicial complexes T0(θ) constructed via pairwise coactivity selection (Method I) for four different threshold values. (A) The appearance rates of the maximal simplexes, arranged according to their dimension, demon- strate remarkably tight, graph-like distribution. The color of the dots corresponds to the dimension of the simplexes, as indicated by the colorbar on the right. (B) Spatial distribution of the dimensionalities of the selected simplexes. (C) Spatial distribution of the appearance rates of the selected simplexes. (D) The histograms of the lapse times, fit to double exponential distribution, and the value of the fitted distribution's parameter β. (E) Spatial projections of the 2D skeletons of T0(θ). Data for all panels is computed for a specific place field map for illustrative purposes. 20 FIG. S6: The selected simplicial complexes T0(n0) constructed via the closed-neighbor selection algorithm (Method II) for four different values of n0. (A) The appearance rates of simplexes, arranged according to their dimensions, color-coded as indicated by the colorbar on the right. (B) Spatial distribution of the dimensionalities of the selected simplexes. (C) Spatial distribution of the appearance rates of the selected simplexes. (D) The histograms of the lapse times, fit to double exponential distribution, and the value of the fitted distribution's parameter β. (E) Spatial projections of the 2D skeletons of T0(n0). VIII. SUPPLEMENTARY TABLES 21 θ = 0.07 Hz θ = 0.05 Hz θ = 0.04 Hz θ = 0.1 Hz 1 139 509 569 1 171 393 410 2 198 324 214 5 131 108 15 1 126 409 480 1 175 353 286 1 203 247 144 7 147 71 17 1 156 580 853 2 211 518 571 2 233 330 419 6 139 152 41 1 135 407 551 1 176 369 333 1 201 296 152 7 139 50 7 1 149 422 585 2 172 342 453 2 185 312 320 13 122 107 73 1 130 485 775 1 186 447 392 1 190 335 250 4 164 84 27 1 106 459 655 1 202 352 603 1 177 395 353 8 179 117 25 1 120 519 787 1 194 397 449 1 215 333 237 8 160 109 63 1 131 441 757 1 165 459 499 1 203 339 197 8 149 65 15 1 160 364 575 1 226 313 368 1 175 270 222 8 117 92 14 TABLE S1: Topological signature of the complex selected by simplex rate thresholding. Each cell contains a list of four Betti numbers, (b0, b1, b2, b3). For the low rate, fσ = 0.04 Hz, T0(θ) encodes the correct spatial connectedness of the environment (b0 = 1). For intermediate rates, 0.05 ≤ fσ ≤ 0.07 Hz, T0(θ) may occasionally break into two pieces (b0 = 2) and for fσ ≥ 0.10 Hz and higher, T0(θ) fragments into multiple components. In higher dimensions D ≥ 1, T0(θ) contains over a hundred noncontractible topological loops. A. Original B. Corrected θ = 0.05 θ = 0.07 θ = 0.1 θ = 0.14 θ = 0.05 θ = 0.07 θ = 0.1 θ = 0.14 1 1 0 0 1 0 0 0 3 0 0 0 1 1 0 0 1 1 0 0 1 0 0 0 2 0 0 0 1 2 0 0 1 1 0 0 1 0 0 0 1 0 0 0 1 2 0 0 1 1 0 0 1 0 0 0 1 0 0 0 1 1 0 0 1 1 0 0 1 0 0 0 3 0 0 0 1 1 0 0 1 1 0 0 1 0 0 0 1 0 0 0 1 3 0 0 1 1 0 0 2 0 0 0 3 0 0 0 2 1 0 0 1 1 0 0 1 0 0 1 2 0 0 0 1 2 0 0 1 1 1 0 1 1 0 0 4 0 0 0 1 3 0 0 1 1 0 0 2 0 0 0 3 0 0 0 1 3 0 0 2 1 0 0 5 1 0 0 3 0 0 0 1 5 0 0 7 2 0 0 7 1 0 0 2 1 0 0 6 2 0 0 7 2 0 0 1 2 0 0 7 7 0 0 7 1 0 0 2 3 0 0 13 1 0 0 8 1 0 0 1 0 0 0 4 6 0 0 11 0 0 0 1 5 0 0 8 1 0 0 7 5 0 0 1 2 0 0 8 2 0 0 9 1 0 0 1 4 0 0 8 2 0 0 9 2 0 0 1 5 0 0 8 2 0 0 7 6 0 0 1 1 0 0 1 1 0 0 1 1 0 0 1 1 0 0 1 1 0 0 1 1 0 0 1 1 0 0 1 1 0 0 1 1 0 0 1 1 0 0 TABLE S2: Topological signature of the complex selected by link-rate thresholding. (A) For the low rate fσ = 0.05 Hz, T0(θ) occasionally produces the correct topological signature (b0 = b1 = 1, bn>1 = 0, shown in bold). For intermediate rates fσ ∼ 0.07 Hz, T0(θ) may occasionally break into two pieces (b0 = 2) and produce extra noncontractible loops in 1D. For high thresholds, fσ ≥ 0.10 Hz, T0(θ) fragments into multiple components. However, the connectivity of T0(θ) in higher dimensions, D ≥ 2, is correct for all cases, which implies that T0(θ) contracts into 2D. (B) After applying the correction algorithms, the selected complexes acquire correct topological signature for fσ ≤ 0.07 for all maps. The corresponding learning times Tmin are listed in Suppl. Table 4. A. Original B. Corrected 22 n0 = 2 n0 = 5 29 0 1 31 0 0 1 11 0 0 1 4 1 0 29 0 1 30 0 0 1 16 0 0 1 3 0 0 32 0 1 28 0 0 1 19 0 0 1 3 0 0 31 0 1 31 0 0 1 14 0 0 1 2 0 0 31 0 2 34 0 0 1 20 0 0 1 4 0 0 34 0 1 26 0 0 1 12 0 0 1 2 0 0 28 0 1 41 0 0 1 11 0 0 1 2 0 0 28 0 1 40 0 0 1 15 0 0 1 1 1 0 31 0 1 32 0 0 1 17 0 0 1 4 0 0 43 0 1 33 0 0 1 16 4 0 1 4 0 0 n0 = 7 n0 = 12 n0 = 2 n0 = 5 n0 = 7 n0 = 12 1 1 3 0 1 1 1 0 1 1 1 1 1 2 0 0 1 1 1 0 1 1 1 0 1 2 0 0 1 1 0 1 1 1 1 0 1 4 0 0 1 1 1 0 1 1 1 0 1 2 0 0 1 1 1 0 1 2 2 1 1 5 0 0 1 1 1 0 1 2 0 1 1 2 0 0 1 2 1 0 1 2 0 0 1 1 0 0 1 1 2 0 1 1 0 2 1 2 0 0 1 3 1 0 1 1 0 0 1 2 0 0 1 1 5 0 1 2 0 0 1 2 1 1 1 4 1 3 1 3 1 2 1 4 1 3 1 3 1 3 TABLE S3: Topological signatures of the complexes selected by the neighbor-selection algorithm. (A) If but one pair of closest vertexes is selected n0 = 2, T0(n0) breaks into multiple components. For n0 ≥ 5, T0(n0) has only one component, but path connectivity is compromised (b0 = 1, b1 (cid:29) 1). In higher dimensions, T0(n0) is contractible, bn>1 = 0. (B) After applying the correction algorithms, the selected complexes for almost all maps acquire correct topological signature in 1D and 2D (shown in boldface) for n0 ≥ 7. The corresponding learning times, Tmin, are listed in Suppl. Table 4. 7 8 9 5 6 2 3 4 T0 1 10 mean/std 4.4 2.7 2.3 2.7 2.8 2.7 2.1 3.8 10.7 2.7 3.7/2.5 T0(θ = .05) 3.8 2.7 1.9 2.7 3.6 3.8 3.8 3.8 2.5 3.7 3.2/0.7 T0(θ = .07) 2.1 3.8 3.8 2.8 2.8 2.3 2.4 3.7 2.5 3.8 3.0/0.7 T0(n0 = 7) 1.9 4.6 2.4 2.7 2.8 3.8 ∞ 3.8 ∞ 3.8 3.2/0.9 T0(n0 = 12) 3.8 2.7 1.9 2.7 ∞ ∞ ∞ 3.8 3.8 ∞ 3.1/0.8 TABLE S4: The learning times, Tmin (in minutes) computed for the selected simplicial complexes T0(θ) and T0(n0) are similar to the learning times computed via the full temporal nerve complex T . Thus, the information about the topological structure of the environment emerges from the cell assembly activity as fast as from the entire pool of place cell coactivities. However, note that the complex T0(n0) sometimes fails to produce a finite leaning time; non-convergent cases are marked by ∞. IX. SUPPLEMENTARY MOVIES ILLUSTRATING THE FIRST OF THE TEN TESTED MAPS Suppl. Movie 1. The grey dots represent centers of the place fields, viewed from above, similar to Figure 3B. The centers of the place fields that correspond to the coactive place cells are shown in red. The resulting activity packet moves in the environment following the simulated rat's trajectory. Suppl. Movie 2. Selection of the cell assemblies by Method I (θ = 100) and assigning readout neurons to the cell assemblies. Suppl. Movie 3. A side projection view of the activity packet propagating in the cell assembly network, selected via Method I. To emphasize that the coactive place cell combinations comprise a cell assembly complex T0(θ), the corresponding place field centers are schematically connected to the readout neurons. Suppl. Movie 4. The same system shown in the same projection as the Figure 3B and Suppl. Movie 1.
1812.11937
1
1812
2018-12-31T18:18:46
Two "correlation games" for a nonlinear network with Hebbian excitatory neurons and anti-Hebbian inhibitory neurons
[ "q-bio.NC", "cs.NE" ]
A companion paper introduces a nonlinear network with Hebbian excitatory (E) neurons that are reciprocally coupled with anti-Hebbian inhibitory (I) neurons and also receive Hebbian feedforward excitation from sensory (S) afferents. The present paper derives the network from two normative principles that are mathematically equivalent but conceptually different. The first principle formulates unsupervised learning as a constrained optimization problem: maximization of S-E correlations subject to a copositivity constraint on E-E correlations. A combination of Legendre and Lagrangian duality yields a zero-sum continuous game between excitatory and inhibitory connections that is solved by the neural network. The second principle defines a zero-sum game between E and I cells. E cells want to maximize S-E correlations and minimize E-I correlations, while I cells want to maximize I-E correlations and minimize power. The conflict between I and E objectives effectively forces the E cells to decorrelate from each other, although only incompletely. Legendre duality yields the neural network.
q-bio.NC
q-bio
Two "correlation games" for a nonlinear network with Hebbian excitatory neurons and anti-Hebbian inhibitory neurons Neuroscience Institute and Computer Science Dept. H. Sebastian Seung Princeton University Princeton, NJ 08544 June 13, 2021 Abstract A companion paper introduces a nonlinear network with Hebbian excitatory (E) neurons that are reciprocally coupled with anti-Hebbian inhibitory (I) neurons and also receive Hebbian feedforward excitation from sensory (S) afferents. The present paper derives the network from two normative principles that are mathe- matically equivalent but conceptually different. The first principle formulates un- supervised learning as a constrained optimization problem: maximization of S−E correlations subject to a copositivity constraint on E − E correlations. A com- bination of Legendre and Lagrangian duality yields a zero-sum continuous game between excitatory and inhibitory connections that is solved by the neural network. The second principle defines a zero-sum game between E and I cells. E cells want to maximize S−E correlations and minimize E −I correlations, while I cells want to maximize I − E correlations and minimize power. The conflict between I and E objectives effectively forces the E cells to decorrelate from each other, although only incompletely. Legendre duality yields the neural network. A companion paper [Seung, 2018] introduces a nonlinear neural network for unsuper- vised learning in which a population of excitatory (E) neurons is reciprocally connected with a population of inhibitory (I) neurons (Fig. 1, left). The E neurons also receive feedforward excitation from a population of sensory (S) afferents. The reciprocal E − I connections allow the E neurons to inhibit each other through disynaptic E → I → E pathways mediated by I neurons. This network motif will be called "disynaptic re- current inhibition," or just "disynaptic inhibition." Excitatory connections (S → E and E → I) are modified by Hebbian plasticity, and inhibitory connections (I → E) by anti- Hebbian plasticity. The companion paper investigates the computational properties of the neural net- work through mathematical analysis and numerical simulations. It would be helpful to attain a deeper understanding of the network as arising from some normative principle of unsupervised learning. As emphasized by Pehlevan and Chklovskii [2015], the long 1 Figure 1: Disynaptic inhibition versus direct all-to-all inhibition. Green arrows indi- cate excitatory connections, which change via Hebbian plasticity. Red arrows indicate inhibitory connections (note minus sign), which change via anti-Hebbian plasticity. (Left) E neurons are reciprocally coupled with I neurons, and receive input from S af- ferents. E → I connections (A) and I → E connections (−A(cid:62)) have equal but opposite strengths. (Right) The E and I populations collapse to a single population of neurons directly connected by all-to-all inhibition (−L). history of neural network models based on Hebbian and anti-Hebbian plasticity has primarily relied on numerical simulations to demonstrate interesting self-organizing behaviors, and lacked normative principles that can be mathematically formalized as optimizations of some objective function. Normative principles are not only helpful for understanding, but also have the practical consequence of suggesting optimization algorithms other than neural networks, which could be useful in certain settings. The companion paper takes one step towards a normative principle by "deriving" the network with disynaptic inhibition (Fig. 1, left) as an approximation to a network with all-to-all inhibition (Fig. 1, right). The latter network contains a single population of neurons that directly inhibit each other and also receive feedforward excitation from sensory afferents, as in the Seung and Zung [2017] variant of the Földiák [1990] model. The derivation involves a factorized approximation of a Lagrange multiplier matrix, which is not fully justified in the companion paper. The present paper is intended to provide the missing theoretical foundation by deriving the network from two normative principles that are mathematically equivalent but suggest different interpretations. The first normative principle is a modification of a principle previously used by Seung and Zung [2017] to derive the network with all-to-all inhibition. They defined unsupervised learning as the maximization of input-output correlations subject to a bound constraint on output-output correlations. The bound constraint can be regarded as enforcing a decorrelated output representation. The principle led to a zero-sum "correlation game" between excitation and all-to-all inhibition in a nonlinear network that was a variant of the Földiák [1990] model.1 The inhibitory connections of the 1Zylberberg et al. [2011] claimed that the Földiák [1990] model learns by fitting a linear generative model to sensory stimuli. The claim is only an approximation at best, and invalid at worst. It is conceptually very different from Seung and Zung [2017] and Pehlevan et al. [2018], who view excitation and inhibition as players in a zero-sum game or minimax problem. According to the linear generative model idea, excitation 2 u1u2u3x1x2x3y1y2WA -- ATS cellsE cellsI cellsu1u2u3x1x2x3W -- LS cells network were conjugate to output-output correlations via Lagrangian duality. The ex- citatory connections were conjugate to input-output correlations via Legendre duality. A similar approach was taken by Pehlevan et al. [2018], who derived linear networks performing variants of principal component analysis from the similarity matching prin- ciple. For the disynaptic inhibition network, the first normative principle is a constrained optimization problem: maximization of input-output correlations subject to a coposi- tivity constraint on output-output correlations. The constraint can be regarded as forc- ing an imperfect form of decorrelation. The principle leads to a zero-sum game solved by the disynaptic inhibition network. The S → E connections are conjugate to the input-output correlations via Legendre duality. The I − E connections are Lagrange multipliers enforcing the copositivity constraint. The first normative principle is based on the activities of the E cells only; the activi- ties of the I cells do not appear until after the duality transforms. The second normative principle is a zero-sum game between E and I cells. E cells want to maximize S− E correlations and minimize E − I correlations. I cells want to maximize E − I correla- tions and minimize power. There is some conflict inherent in the fact that I and E cells want to drive E − I correlations in opposite directions. The compromise is that E cells tend to decorrelate from each other. Both excitatory and inhibitory connections arise from Legendre duality. A theatrical metaphor describes the difference between the normative principles. The first normative principle places the E cells center stage while the I cells lurk back- stage. In the second normative principle, E and I cells are both on stage; E cells are leading actors while I cells are supporting actors. No connections appear in either normative principle; the network only emerges after duality transforms. Both normative principles are mathematically equivalent to each other, but they invite different viewpoints. The first normative principle is simple because it contains E cell activity only, but complex because copositivity is not a terribly intuitive idea. The second normative principle is complex because it contains both E and I cell activity, but simple because there is no need for copositivity. The second normative principle is similar to two algorithms for dimensionality reduction (hard-thresholding of covariance eigenvalues and equalizing thresholded co- variance eigenvalues) defined by Pehlevan and Chklovskii [2015]. The neural network implementation of the algorithms involved separate populations of principal neurons and interneurons, but the neurons did not obey Dale's Law and so could not be inter- preted as excitatory and inhibitory neurons. and inhibition minimize reconstruction error. 3 1 Network model with disynaptic inhibition (cid:33)(cid:35)+ The disynaptic inhibition network (Fig. 1, left) has the activity dynamics, (cid:34) (cid:32) m xi := (1− dt)xi + dt λ−1 i Wiaua − r ∑ α=1 ∑ a=1 yα = n ∑ i=1 Aαixi yαAαi (1) (2) Here dt is a step size parameter, which can be set at a small constant value or adjusted adaptively. The activation function [z]+ = max{z,0} is half-wave rectification. After the activities converge to a steady state, update the connection matrices via ∆Wia ∝ xiua − γWia − κ ∑ ∆Aα j ∝ yαx j −(cid:0)q2 − p2(cid:1)Aα j − p2∑ Wib b i Aαi (3) (4) where γ > 0, κ > 0, and q2 > p2 . After the updates (3) and (4), any negative elements of W and A are zeroed to maintain nonnegativity. The divisive factor λi > 0 in Eq. (1) is updated via (5) Intuitions behind the model definitions are explained in the companion paper [Seung, 2018]. The goal of the present paper is show how the network can be interpreted as a method of solving a zero-sum game. ∆λi ∝ x2 i − q2 2 Correlation game between connections 2.1 Formulation as constrained optimization The first normative principle concerns transformation of a sequence of input vectors u(1), . . . ,u(T ) into a sequence of output vectors x(1), . . . ,x(T ). Both input and output are assumed nonnegative. Define the input matrix U = [u(1), . . . ,u (T )] as the matrix containing input vectors u(t) as its columns. The element Uat is the ath component of u(t). Similarly, define the output matrix X = [x (1) , . . . ,x (T )] as containing output vectors x(t) as its columns. Define the output-input correlation matrix is XU(cid:62) T = 1 T T ∑ t=1 (cid:62) x (t)u (t) Its ia element is the time average of xiua, or (cid:104)xiua(cid:105). Similarly, define the output-output correlation matrix XX(cid:62) T = 1 T T ∑ t=1 (cid:62) x (t)x (t) 4 Its i j element is the time average of xix j, or (cid:104)xix j(cid:105). Note that "correlation matrix" is used to mean second moment matrix rather than covariance matrix. In other words, the correlation matrix does not involve subtraction of mean values. This is natural for sparse nonnegative variables, but covariance matrices may be substituted in other settings. Problem 1 (Constrained optimization). Define the goal of unsupervised learning as the constrained optimization subject to copositivity of D− XX(cid:62) max X≥0 (6) where D is a fixed matrix and Φ∗ is a scalar-valued function that is assumed monotone nondecreasing as a function of every element of its matrix-valued argument. T T Φ∗(cid:18) XU(cid:62) (cid:19) Monotonicity is an important assumption because it allows us to interpret the ob- jective of Eq. (6) as maximization of input-output correlations. 2.2 Copositivity vs. nonnegativity Seung and Zung [2017] introduced the principle Φ∗(cid:18) XU(cid:62) (cid:19) T max X≥0 subject to nonnegativity of D− XX(cid:62) T (7) which differs from Eq. (6) only by the substitution of "nonnegativity" for "copos- itivity." (Here nonnegativity of a matrix is defined to mean nonnegativity of all its elements.) While the formalisms here are valid for arbitrary D, a convenient choice is to set diagonal elements of D to be q2 and off-diagonal elements of D to be p2, (cid:40) Di j = q2, p2, i = j, i (cid:54)= j (8) If p is much smaller than q, the nonnegativity constraint (cid:104)xix j(cid:105) ≤ Di j in Eq. (7) amounts to decorrelation. A symmetric matrix S is said to be copositive when v(cid:62)Sv ≥ 0 for every nonnegative vector v ≥ 0. This constraint is analogous to positive semidefiniteness but is more complex because it cannot be reduced to a single eigenvalue constraint. Hahnloser et al. [2003] give sufficient and necessary conditions for copositivity involving eigenvalues of submatrices. Nonnegativity of S is a sufficient condition for copositivity of S, but it is not a neces- sary condition. In particular, copositivity of S−XX(cid:62)/T does not require nonnegativity, so a solution of Problem 1 may have (cid:104)xix j(cid:105) > Di j for some i and j. A necessary condition for copositivity of S is nonnegativity of its diagonal ele- ments, since e(cid:62) i Sei < 0 if Sii < 0 where e1, . . . ,en denotes the standard basis for Rn. In particular copositivity of S = D− XX(cid:62)/T requires that (cid:104)x2 i (cid:105) ≤ Dii for all i. These inequalities will be called "power constraints," because they limit the power in the outputs. 5 If either of the diagonal elements Sii and S j j vanish, then a necessary condition for copositivity is nonnegativity of the off-diagonal element Si j. Therefore (cid:104)xix j(cid:105) may exceed Di j in a solution of Problem 1 only if the power constraints for i and j are not saturated. 2.3 Correlation game from Legendre-Lagrangian duality The copositivity constraint in Eq. (6) can be enforced by introducing Lagrange multi- pliers A and Λ, (cid:26) Φ∗(cid:18) XU(cid:62) (cid:19) T (cid:19) (cid:18) D− XX(cid:62) T + 1 2 TrA AT + 1 2 TrΛ (cid:19)(cid:27) (cid:18) D− XX(cid:62) T (9) max X≥0 min A,Λ≥0 The Lagrange multiplier A is a nonnegative r × n matrix. The outer maximum must choose X so that D− XX(cid:62)/T is copositive because otherwise the minimum with re- spect to A is −∞. The Lagrange multiplier Λ = diag{λ1, . . . ,λn} is a nonnegative diagonal matrix. The outer maximum must choose X so that the diagonal elements of D− XX(cid:62)/T are nonnegative because otherwise the minimum with respect to Λ is −∞. As mentioned above, copositivity of D − XX(cid:62)/T by itself already implies that the diagonal elements are nonnegative. It follows that the Lagrange multiplier Λ is redundant for the primal problem, though it does affect the dual problem. Similarly, adding extra rows to the Lagrange multiplier A does not change the primal problem. For enforcing the copositivity constraint, it would be sufficient for A to be 1× n. However, making r > 1 does affect the dual problem. Problem 2 (Game between cells and connections). Switching the order of min and max in Eq. (1) yields the dual problem, (cid:26) Φ∗(cid:18) XU(cid:62) (cid:19) T (cid:18) D− XX(cid:62) T (cid:19)(cid:16) + 1 2 Tr min A,Λ≥0 max X≥0 (cid:17)(cid:27) A(cid:62)A + Λ (10) This is an upper bound for Eq. (9) by the minimax inequality. At this point, it is convenient to define the objective function Φ∗ as the convex conjugate (Legendre-Fenchel transform) of a function Φ, (cid:40) ∑ ia (cid:41) Φ∗(C) = max W≥0 WiaCia − Φ (W ) . (11) (11) guarantees that Φ∗(C) is monotone The nonnegativity constraint on W in Eq. nondecreasing as a function of every element of C. The function Φ can be interpreted as a regularizer or prior for the weight matrix W . With Legendre duality, a maximization with respect to W is implicit in Eq. (11). Switching the order of W and X maximizations yields the following equivalent prob- lem. 6 Problem 3 (Game between connections). The Lagrangian dual of the constrained op- timization in Problem 1 is min A,Λ≥0 max W≥0 R(W,A(cid:62)A + Λ) with payoff function defined by (cid:26) 1 T R(W,L) = max X≥0 TrW(cid:62)XU(cid:62) − Φ(W ) + (cid:19) (cid:18) D− XX(cid:62) T (cid:27) L 1 2 Tr (12) (13) The min-max problem can be interpreted as a zero-sum game between W on the one hand and A and Λ on the other. Problem 3 is closely related to the correlation game previously introduced by Seung and Zung [2017], (14) Problem 3 constrains L = A(cid:62)A + Λ for some nonnegative A and Λ, so it is an upper bound for Eq. (14). This is the mathematical interpretation of choosing a parametrized form A(cid:62)A + Λ for the Lagrange multiplier L, as was done by Seung [2018]. R(W,L) max W≥0 min L≥0 The network model of Section 1 follows by setting Φ(W ) = γ 2 ∑ ia W 2 ia + κ 2 ∑ i (cid:19)2 (cid:18) ∑ a Wia in Eq. (13) and applying online projected gradient ascent to perform the maximizations in Eq. (12) and online projected gradient descent to perform the minimizations. For a more general choice of Φ(W ), Eq. (3) should be replaced by ∆Wia ∝ xiua − ∂Φ ∂Wia 3 Correlation game between cells The second normative principle concerns transformation of a sequence of nonnega- tive input vectors u(1), . . . ,u(T ) into two sequences of nonnegative output vectors x(1), . . . ,x(T ) and y(1), . . . ,y(T ). Define the input matrix U = [u(1), . . . ,u (T )] and the two output matrices X = [x (1) , . . . ,x (T )] and Y = [y (1) , . . . ,y (T )]. Problem 4 (Game between cells). Define the goal of unsupervised learning as the zero-sum game between X and Y (cid:26) Φ∗(cid:18) XU(cid:62) (cid:19) − Ψ∗(cid:18)Y X(cid:62) (cid:19) (cid:27) 1 2 YY(cid:62) T min Y≥0 max X≥0 (15) where Φ∗ and Ψ∗ are scalar-valued functions assumed monotone nondecreasing as a function of every element of their matrix-valued arguments. Tr + T T 7 Note that only nonnegativity constraints remain in Problem 4; the copositivity con- straint of Problem 1 is completely hidden. This correlation game can be interpreted as follows. The E cells would like to maximize S− E correlations (make Φ∗ large) and minimize E −I correlations (make Ψ∗ small). The I cells would like to maximize I−E correlations (make Ψ∗ large) and minimize power (make TrYY(cid:62)/T small). There is conflict between the E and I cells because E cells would like to minimize E − I cor- relations while I cells would like to maximize them. The compromise is that E cells incompletely decorrelate from each other. Problem 4 is equivalent to Problem 1 given the definition of Ψ∗(C) = max A≥0 as the Legendre transform of (cid:41) AαiCαi − Ψ(A) (cid:40) ∑ iα Ψ(A) = TrADA(cid:62) 1 2 (16) (17) (cid:27) (cid:26) Φ∗(cid:18) XU(cid:62) (cid:19) (cid:26) Φ∗(cid:18) XU(cid:62) T T (cid:19) Proof. Substituting the definitions of Eqs. (16) and (17) into Eq. (15) yields max X≥0 min A,Y≥0 − 1 T TrA(cid:62)Y X(cid:62) + TrADA(cid:62) + 1 2 1 2 Tr YY(cid:62) T This is minimized when Y = AX, attaining the value max X≥0 min A≥0 − 1 2T TrA(cid:62)AXX(cid:62) + 1 2 TrADA(cid:62)(cid:27) This is identical to Eq. (9), except for the omission of the Lagrange multiplier Λ which, as mentioned previously, is redundant in the primal problem. 4 Discussion The second normative principle is also interesting because it can be generalized to include E − E and I − I connections. This will be the subject of future work. Acknowledgments The author is grateful for helpful discussions with J. Zung, C. Pehlevan and D. Chklovskii. The research was supported in part by the Intelligence Advanced Research Projects Activity (IARPA) via DoI/IBC contract number D16PC0005, and by the National In- stitutes of Health via U19 NS104648 and U01 NS090562. 8 References Peter Földiák. Forming sparse representations by local anti-hebbian learning. Biologi- cal cybernetics, 64(2):165 -- 170, 1990. Richard HR Hahnloser, H Sebastian Seung, and Jean-Jacques Slotine. Permitted and forbidden sets in symmetric threshold-linear networks. Neural computation, 15(3): 621 -- 638, 2003. Cengiz Pehlevan and Dmitri Chklovskii. A normative theory of adaptive dimension- ality reduction in neural networks. In Advances in neural information processing systems, pages 2269 -- 2277, 2015. Cengiz Pehlevan, Anirvan M Sengupta, and Dmitri B Chklovskii. Why do similarity matching objectives lead to hebbian/anti-hebbian networks? Neural computation, 30(1):84 -- 124, 2018. H Sebastian Seung. Unsupervised learning by a nonlinear network with hebbian exci- tatory and anti-hebbian inhibitory neurons. arXiv, 2018. H Sebastian Seung and Jonathan Zung. A correlation game for unsupervised learning yields computational interpretations of hebbian excitation, anti-hebbian inhibition, and synapse elimination. arXiv preprint arXiv:1704.00646, 2017. Joel Zylberberg, Jason Timothy Murphy, and Michael Robert DeWeese. A sparse coding model with synaptically local plasticity and spiking neurons can account for the diverse shapes of v1 simple cell receptive fields. PLoS Comput Biol, 7(10): e1002250, 2011. 9
1102.0566
2
1102
2011-07-20T15:57:22
Model Cortical Association Fields Account for the Time Course and Dependence on Target Complexity of Human Contour Perception
[ "q-bio.NC" ]
Can lateral connectivity in the primary visual cortex account for the time dependence and intrinsic task difficulty of human contour detection? To answer this question, we created a synthetic image set that prevents sole reliance on either low-level visual features or high-level context for the detection of target objects. Rendered images consist of smoothly varying, globally aligned contour fragments (amoebas) distributed among groups of randomly rotated fragments (clutter). The time course and accuracy of amoeba detection by humans was measured using a two-alternative forced choice protocol with self-reported confidence and variable image presentation time (20-200 ms), followed by an image mask optimized so as to interrupt visual processing. Measured psychometric functions were well fit by sigmoidal functions with exponential time constants of 30-91 ms, depending on amoeba complexity. Key aspects of the psychophysical experiments were accounted for by a computational network model, in which simulated responses across retinotopic arrays of orientation-selective elements were modulated by cortical association fields, represented as multiplicative kernels computed from the differences in pairwise edge statistics between target and distractor images. Comparing the experimental and the computational results suggests that each iteration of the lateral interactions takes at least 37.5 ms of cortical processing time. Our results provide evidence that cortical association fields between orientation selective elements in early visual areas can account for important temporal and task-dependent aspects of the psychometric curves characterizing human contour perception, with the remaining discrepancies postulated to arise from the influence of higher cortical areas.
q-bio.NC
q-bio
Model Cortical Association Fields Account for the Time Course and Dependence on Target Complexity of Human Contour Perception Vadas Gintautas,1, 2, ∗ Michael I. Ham,3 Benjamin Kunsberg,4 Shawn Barr,4 Steven P. Brumby,5 Craig Rasmussen,4 John S. George,6 Ilya Nemenman,7 Lu´ıs M. A. Bettencourt,1 and Garrett T. Kenyon4, † 1Center for Nonlinear Studies and T-5, Theoretical Division, Los Alamos National Laboratory, Los Alamos, New Mexico, USA 2Physics Department, Chatham University, Pittsburgh, Pennsylvania, USA 3P-21, Physics Division, Los Alamos National Laboratory, Los Alamos, New Mexico, USA 4New Mexico Consortium, Los Alamos, New Mexico, USA 5Space and Remote Sensing Sciences, Los Alamos National Laboratory, Los Alamos, New Mexico, USA 6P-21 Applied Modern Physics (Biological and Quantum Physics), Los Alamos National Laboratory, Los Alamos, New Mexico, USA 7Departments of Physics and Biology and Computational and Life Sciences Initiative, Emory University, Atlanta, Georgia, USA (Dated: November 5, 2018) Abstract Author Summary Can lateral connectivity in the primary visual cortex account for the time dependence and intrinsic task diffi- culty of human contour detection? To answer this ques- tion, we created a synthetic image set that prevents sole reliance on either low-level visual features or high-level context for the detection of target objects. Rendered images consist of smoothly varying, globally aligned con- tour fragments (amoebas) distributed among groups of randomly rotated fragments (clutter). The time course and accuracy of amoeba detection by humans was mea- sured using a two-alternative forced choice protocol with self-reported confidence and variable image presentation time (20-200 ms), followed by an image mask optimized so as to interrupt visual processing. Measured psychome- tric functions were well fit by sigmoidal functions with exponential time constants of 30-91 ms, depending on amoeba complexity. Key aspects of the psychophysical experiments were accounted for by a computational net- work model, in which simulated responses across retino- topic arrays of orientation-selective elements were mod- ulated by cortical association fields, represented as mul- tiplicative kernels computed from the differences in pair- wise edge statistics between target and distractor im- ages. Comparing the experimental and the computa- tional results suggests that each iteration of the lateral interactions takes at least 37.5 ms of cortical processing time. Our results provide evidence that cortical asso- ciation fields between orientation selective elements in early visual areas can account for important temporal and task-dependent aspects of the psychometric curves characterizing human contour perception, with the re- maining discrepancies postulated to arise from the influ- ence of higher cortical areas. ∗Electronic address: [email protected] †Electronic address: [email protected] Current computer vision algorithms reproducing the feed-forward features of the primate visual pathway still fall far behind the capabilities of human subjects in de- tecting objects in cluttered backgrounds. Here we inves- tigate the possibility that recurrent lateral interactions, long hypothesized to form cortical association fields, can account for the dependence of object detection accuracy on shape complexity and image exposure time. Corti- cal association fields are thought to aid object detection by reinforcing global image features that cannot easily be detected by single neurons in feed-forward models. Our implementation uses the spatial arrangement, rela- tive orientation, and continuity of putative contour el- ements to compute the lateral contextual support. We designed synthetic images that allowed us to control ob- ject shape and background clutter while eliminating un- intentional cues to the presence of an otherwise hidden target. In contrast, real objects can vary uncontrollably in shape, are camouflaged to different degrees by back- ground clutter, and are often associated with non-shape cues, making results using natural image sets difficult to interpret. Our computational model of cortical associa- tion fields matches many aspects of the time course and object detection accuracy of human subjects on statis- tically identical synthetic image sets. This implies that lateral interactions may selectively reinforce smooth ob- ject global boundaries. Introduction The perception of closed contours is fundamental to object recognition, as revealed by the fact that com- mon object categories can be rapidly detected in black and white line drawings in which all shading and lumi- nance cues have been removed [1]. Cortical association fields, hypothesized to capture spatial correlations be- tween local image features via long-range lateral synap- tic interactions, provide a natural substrate for rapid contour perception [2]. The link between cortical as- sociation fields and contour perception has been inves- tigated through a variety of behavioral, experimental, and theoretical techniques [3 -- 6]. Psychophysical mea- surements reveal that the detection of implicit contours, defined by sequences of Gabor-like elements presented against randomly oriented backgrounds, becomes more difficult as the local curvature increases and as the indi- vidual Gabor elements are spaced further apart or their alignment is randomly perturbed. This dependence on proximity and relative orientation implies that, in early visual areas, cortical association fields are primarily lo- cal and aligned along smooth trajectories [2, 7, 8]. In re- lated studies, collinear Gabor patches have been shown to both increase and decrease the contrast detection thresh- old of a central Gabor patch in a manner that depends on the relative timing, orientation and spatial separation of the flanking elements [9 -- 11], providing further psy- chophysical evidence that lateral influences act at early cortical processing stages, although the contribution of collinear facilitation to contour integration remains con- troversial [12]. In primary visual cortex (V1), electro- physiological recordings indicate that the responses to optimally oriented and positioned stimuli can be facili- tated by flanking stimuli placed outside the classical re- ceptive field center [5, 6, 10, 13], although these effects have also been ascribed to elongated central receptive fields [14, 15] and facilitation has been attributed to in- creases in baseline activity [16]. Nonetheless, collinear facilitation is consistent with anatomical studies indi- cating that orientation columns are laterally connected to surrounding columns with similar orientation prefer- ence [17 -- 19]. Because extensive association fields are present in the primary visual cortex [17 -- 19], lateral interactions may be key to discriminating smooth object boundaries at very fast time scales (of the order of tens of ms), as ob- served in numerous speed of sight psychophysical exper- iments [1, 20 -- 23]. Correspondingly, theoretical models have proposed that V1 cortical association fields can be described mathematically on the basis of cocircularity, and that relaxation dynamics based on cocircular asso- ciation fields can extract global contours by suppress- ing local variation [24]. Such models are qualitatively consistent with human judgments as to whether pairs of short line segments belong to the same or separate con- tours, with human judgments closely following the pair- wise statistics of edge segments extracted from natural scenes [25]. Further, model cortical association fields, when used to detect implicit contours, can predict key aspects of human psychophysics, particularly the mea- sured dependence on the density of foreground elements relative to background elements [8, 26]. In this paper, we extend the above studies by inves- tigating whether model cortical association fields can account not only for dependence of contour perception on intrinsic task difficulty, a relationship that has been previously explored [8, 26], but also for the detailed 2 time course of human contour detection, an aspect that has heretofore not been modeled explicitly, although the time-dependent influence of lateral interactions has been determined for several theoretical models [27, 28]. In this work, we employ multiplicative relaxational dynamics to estimate the time course of contour detection from a com- putational model employing optimized kernels. Model re- sults are then compared to speed-of-sight measurements from human subjects performing the same contour de- tection task. To obtain optimized cortical association fields, we design lateral connectivity patterns using a novel method that exploits the global statistical prop- erties of salient contours relative to background clut- ter. Our procedure, which can be generalized beyond the present application, can be summarized as follows. We begin by generating a large training corpus, di- vided into target and distractor images, from which we obtain estimates of the pairwise co-occurence probabil- ity of oriented edges conditioned on the presence or ab- sence of globally salient contours. From the difference in these two probability distributions, we construct Object- Distractor Difference (ODD) kernels, which are then con- volved with every edge feature to obtain the lateral con- textual support at each location and orientation across the entire image. Edge features that receive substantial contextual support from the surrounding edges are pre- served, indicating they are likely to belong to a globally salient contour, whereas edge features receiving minimal contextual support are suppressed, indicating they are more likely to be part of the background clutter. The lateral contextual support is applied in a multiplicative fashion, so as to prevent the appearance of illusory edges, and the process is iterated several times, mimicking the exchange of information along horizontal connections in the primary visual cortex. Our method is thus intended to capture the essential computational elements of cor- tical association fields that are hypothesized to mediate the pop-out of salient contours against cluttered back- grounds. To obtain a large number of training images and to better isolate the role of cortical association fields linking low-level visual features, we employ abstract computer- generated shapes consisting of short, smooth contour seg- ments that could either be globally aligned to form wig- gly, nearly closed objects (amoebas), or else randomly rotated to provide a background of locally indistinguish- able contour fragments (clutter). Amoeba targets lack specific semantic content, presumably reducing the in- fluence of high level cortical areas, such as IT. However, our computer-generated images would not be expected to eliminate the contribution to contour perception from ex- trastriate areas [29 -- 32]. Thus, our model of lateral inter- actions between orientation-selective neurons is designed to account for just one of several cortical mechanisms that likely contribute to contour perception. Our amoeba/no-amoeba image set differs from stim- uli used in previous psychophysical experiments that employed sequences of Gabor-like elements to repre- sent salient contours against randomly oriented back- grounds [2, 7, 8]. An advantage of contours represented by random Gabor fields is that the target and distrac- tor Gabor elements can be distributed at approximately equal densities, thereby precluding the use of local den- sity operators as surrogates for global contour percep- tion [2]. However, our amoeba/no-amoeba image set is more akin to the natural image sets used in previ- ous speed-of-sight object detection tasks [33], particu- larly with respect to studies employing line drawings de- rived from natural scenes [1]. Humans can detect closed contours, whether defined by aligned Gabor elements or by continuous line fragments, in less than 200 ms [1, 20], which is shorter than the mean interval between sac- cadic eye movements [34], thus mitigating the contribu- tion from visual search. Like Gabor defined contours, our amoeba/no-amoeba image set implements a pop-out detection task involving readily perceived target shapes whose complexity can be controlled parametrically. To benchmark the accuracy and the time course of the ODD kernel-based procedure applied to the amoeba/no- amoeba task, we compare our model results to the per- formance of human subjects on a 2AFC speed-of-sight task in which amoeba/no-amoeba images are presented very briefly side by side, followed by a mask designed to limit the time the visual system is able to process the sensory input [1, 20 -- 23]. Since it takes an estimated 100− 300 ms for activation to spread through the ventral stream of the visual cortex [21], an effective mask pre- sented within this time frame can potentially degrade ob- ject detection performance by interfering with the neural processing mechanisms underlying recognition [22, 35]. By plotting task performance as a function of the stimu- lus onset asynchrony (SOA) -- the interval between image and mask presentation onsets -- the resulting psychometric curves are hypothesized to estimate the neural processing time required to reach a given level of classification ac- curacy. Amoeba targets of low to moderate complexity were found to reliably pop-out against the background clutter, allowing subjects to achieve near perfect perfor- mance at SOAs less than 250 ms, even when followed by an optimized mask consisting of rotated versions of the target and distractor images [20]. Our model cortical as- sociation fields were able to account for the dependence of human performance on amoeba complexity as well as for aspects of the time course of contour perception as measured by the improvement in human performance with increasing SOA. Thus, we present the first network- level computational model to simultaneously account for spatial and temporal aspects of contour perception, as measured in human subjects performing the same con- tour detection task. Aspects of the experimental data for which our model fails to account, particularly data showing that human subjects require longer processing times to detect more complex targets, may indicate the possible involvement of extrastriate areas, which may be essential for the perception of more complex shapes. 3 FIG. 1: Examples of targets and distractors from the amoeba/no-amoeba image set for different K. From top to bottom: K = 2, 4, 6, 8. Left column: Targets; amoeba complexity increases with increasing numbers of ra- dial frequencies. Clutter was constructed by randomly ro- tating groups of amoeba contour fragments. Right column: Distractors; only clutter fragments are present. Results To investigate low-level cortical mechanisms for detect- ing smooth, closed contours presented against cluttered backgrounds with statistically similar low-level features, we designed an amoeba/no-amoeba detection task using a novel set of synthetic images (Figure 1). Amoebas are radial frequency patterns [36] constructed via superposi- tion of periodic functions described by a discrete set of radial frequencies around a circle. In addition, we added clutter objects, or distractors, that were locally indis- tinguishable from targets. Both targets and distractors were composed of short contour fragments, thus elimi- nating unambiguous indicators of target presence or ab- sence, such as total line length, the presence of line end- points, and the existence of short gaps between opposed line segments. To keep the bounding contours smooth, only the lowest K radial frequencies were included in the linear superposition used to construct amoeba targets. To span the maximum range of contour shapes and sizes, the amplitude and phase of each radial frequency com- ponent was chosen randomly, under the restriction that the minimum and maximum diameters could not exceed lower and upper limits. When only 2 radial frequencies were included in the superposition, the resulting amoe- bas were very smooth. As more radial frequencies were included, the contours became more complex. Thus, K, the number of radial frequencies included in the superpo- sition, provided a control parameter for adjusting target complexity. Figure 1 shows target and distractor images generated using different values of K. lateral Human subjects are able to infer whether a two iso- lated line segments extracted from a natural scene are from the same or from separate contours using only distance, direction and relative orientation of the two segments as cues [25, 37]. The performance of human subjects is well predicted by differences in the empiri- cally calculated co-occurrence statistics of short line seg- ments drawn from either the same or from different con- tours. To explore the ability of cortical association fields to account for the perception of smooth contours, we developed a network-level computational model of lat- eral interactions between orientation-selective elements governed by sigmoidal (piecewise linear) input/output synaptic transfer functions. To model inter- actions, we constructed "Object-Distractor Difference (ODD) kernels" for the amoeba/no-amoeba task by com- puting coactivation statistics for the responses of pairs of orientation-selective filter elements, compiled separately for target and distractor images (Figure 2). Because the amoeba/no-amoeba image set was translationally invari- ant and isotropic, the central filter element may without loss of generality be shifted and rotated to a canonical po- sition and orientation. Thus the canonical ODD kernel was defined relative to filter elements at the origin with orientation π/16 (to mitigate aliasing effects). Filter el- ements located away from the origin can be accounted for by a trivial translation. To account for filter elements with different orientations, separate ODD kernels were computed for 8 orientations then rotated to a common orientation and averaged to produce a canonical ODD kernel. The canonical kernel was then rotated in steps between 0 and π (offset by π/16) and then interpolated to Cartesian x−y axes by rounding to the nearest integer coordinates. The resulting ODD kernels were generally consistent with the predictions of cocircular constructions [24], ex- cept that support was mostly limited to line elements lying along low curvature contours, which follows natu- rally from the prevalence of low curvatures in our amoeba training set. Curiously, the largest differences in the coactivation statistics occur close to the center of the kernel, where 4 targets and distractors are presumably most similar. However, even at short distances, amoeba segments are still more likely to be aligned than clutter elements. Moreover, nearby pairs occur much more frequently than more distant pairs, amplifying their contribution to the difference map. Since, by design, the individual clutter fragments were locally indistinguishable from the tar- get fragments, co-occurrence statistics of oriented frag- ments were necessary to solve the amoeba/no-amoeba task. The simplest solution, adopted here, was to fo- cus on pairwise co-occurrences. Notably, in some neural preparations, pairwise interactions have been shown to be sufficient to account for a large fraction of all higher- order correlations [38, 39]. At the retinal stage, target and distractor images were represented as 256 × 256 pixel monochromatic, binary line drawings. At the next stage, corresponding to an early cortical processing area such as V1, a set of filters was used to represent 8 orientations, uniformly-spaced and centered at each pixel, with the axes rotated slightly (by π/16) to mitigate aliasing artifacts. The bottom- up responses of each orientation-selective element were computed via linear convolution using filters composed of a central excitatory subunit flanked by two inhibitory subunits. Each subunit was an elliptical Gaussian with an aspect ratio of 7 : 1, consistent with the aspect ra- tios of V1 simple cell receptive fields measured experi- mentally [40] and similar to values employed in previ- ously published models of V1 responses [41]. Likewise, we estimate that each image pixel subtended a visual an- gle of approximately 0.025◦ (see Methods), so that each orientation-selective element in the model subtended a vi- sual angle of approximately 0.2◦, consistent with physio- logical estimates of V1 receptive field sizes at small eccen- tricities [42]. All subunits had the same total integrated strength (to within a sign), whose magnitude was ad- justed to yield relatively clean representations of the orig- inal image in terms of oriented edges. The synaptic trans- fer function was piecewise-linear with a minimum value of 0.0 and a maximum value of 1.0 and a fixed threshold of 0.5. A finite threshold and saturation level were essen- tial in order to allow non-supported contour fragments to be suppressed while preventing well-supported frag- ments from growing without bound. The precise values used for threshold and saturation were not critical, as responsiveness was controlled independently by adjust- ing the overall integrated strength of the bottom-up and lateral interaction kernels (see Methods). Orientation-selective responses were modulated by 4 successive applications of the multiplicative ODD kernel. Lateral support was first computed via linear convolu- tion of the ODD kernel with the surrounding orientation- selective elements, out to a radius of 32 pixels. Given that images were approximately 7◦ × 7◦ in extent (see Methods), ODD kernels spanned a total visual angle of approximately 1.75 degrees, roughly in correspondence with the estimated visuotopic extent of horizontal pro- jections in V1 [42]. The previous activity of each cell 5 FIG. 2: ODD kernels. Top Row: For a single short line segment oriented approximately horizontally at the center (not drawn), the co-occurrence-based support of other edges at different relative orientations and spatial locations is depicted. Axes were rotated by (180◦/16) from vertical to mitigate aliasing effects. The color of each edge was set proportional to its co- occurrence-based support. The color scale ranges from blue (negative values) to white (zero) to red (positive values). Left panel: Co-occurrence statistics compiled from 40, 000 target images. Center panel: Co-occurrence statistics compiled from 40, 000 distractor images. Right panel: ODD kernel, given by the difference in co-occurrence statistics between target and distractor kernels. Bottom Row: Subfields extracted from the middle of the upper left quadrant (as indicated by black boxes in the top row figures), shown on an expanded scale to better visualize the difference in co-occurrence statistics between target and distractor images. Alignment of edges in target images is mostly cocircular whereas alignment is mostly random in distractor images, accounting for the fine structure in the corresponding section of the ODD kernel. was multiplied by the current lateral support, passed through the piecewise-linear synaptic transfer function, and the process repeated for up to 4 iterations. Contour segments that received insufficient lateral support were thereby suppressed, whereas strongly supported elements were either enhanced or remained maximally activated. When applied to the amoeba/no-amoeba image set, the ODD kernels typically suppressed clutter relative to tar- get segments (Figure 3, left column). When applied in a similar manner to a natural gray- scale image to which a hard Difference-of-Gaussians (DoG) filter has been applied to maximally enhance lo- cal contrast (see Figure 3, right column), ODD-kernels tended to preserve long, smooth lines while suppressing local spatial detail. Although ODD kernels were trained on a narrow set of synthetic images, the results exhibit some generalization to natural images due to the overlap between the cocircularity statistics (see Figure 2) of the synthetic image set and those of natural images. To quantify the ability of the model to discriminate be- tween amoeba/no-amoeba target and distractor images, we used the total activation summed over all orientation- selective elements after k iterations of the ODD kernel. A set of 2, 000 target and distractor images was used for testing; test images were generated independently from the training images. Histograms of the total ac- tivation show increasing separability between target and distractor images as a function of the number of itera- tions (Figure 4). To maximize the range of shapes and sizes spanned by our synthetic targets and distractors, we did not require that the number of ON retinal pixels be constant across images. Rather, the retinal repre- sentations of both target and distractor images encom- passed a broad range of total activity levels, although the two distributions strongly overlapped and there was no evident bias favoring one or the other. At the next processing stage, prior to any lateral interactions, there was likewise little or no bias evident in the bottom-up responses of the orientation-selective elements. Each it- eration of the multiplicative ODD kernel then caused the distributions of total activity for target and distractor images to become more separable, implying correspond- ing improvements in discrimination performance on the amoeba/no-amoeba task. The general principles governing the operation of our model cortical association fields are conceptually straightforward. ODD kernels, which capture differences in the coactivation statistics of edge segments belong- 6 ing to amoebas relative to edge segments belonging to the background clutter, are used to determine the lateral contextual support for individual edge segments in an image. Edge segments receiving sufficiently strong sup- port are preserved, indicating they are likely to be part of an amoeba, whereas edge segments receiving insuffi- cient support are suppressed, indicating they are likely to belong to the background clutter. To assess the ability of the model cortical association fields to account for the time course of human contour perception, we measured the stimulus presentation time required for human subjects to reach a given level of accu- racy on an amoeba/no-amoeba task. The psychophysical experiment was implemented using a speed-of-sight pro- tocol employing a two-alternative forced choice (2AFC) design, with subjects using a slider bar to indicate which of two images, presented side-by-side, contained an amoeba (Figure 5). The distance the bar was dis- placed to the left or to the right was used to indicate confidence, see Methods. To effectively interrupt visual processing at a given SOA, both target and distractor images were replaced by an optimized mask, constructed by combining randomly rotated amoeba and clutter seg- ments [20]. Our optimized masks were designed to render the amoeba targets virtually invisible in the fused target- mask composite. As a measure of human performance on the amoeba/no-amoeba task, we constructed receiver oper- ating characteristic (ROC) curves [43] (Figure 6), using each subject's reported confidence (slider bar location rel- ative to the center position) as a noisy signal for estimat- ing which side, either left or right, contained the target on a given trial. True positives corresponded to trials on which the subject reported the target was on the left (relative to threshold) and the target was actually on the left (relative to threshold). False positives corresponded to trials on which the subject reported the target was on the left whereas the target was actually on the right (relative to threshold). To construct each ROC curve, the confidence scale along the slider bar was divided into 6 discrete threshold values. For each threshold value, a cumulative proportional true positive rate was calculated by considering only those trials as true positives in which the confidence value was above threshold. The cumu- lative proportional false positive rate for each threshold value was calculated similarly. Each threshold value thus contributed one point on the ROC curve, with true pos- itive rate plotted as the ordinate and the false positive rate as the abscissa. The complete set of points were con- nected by straight lines to guide the eye (Figure 6), with a separate ROC curve computed for each combination of SOA and target complexity. ROC curves for quantifying the performance of the model on the amoeba/no-amoeba task were computed similarly, using the difference in total luminance between the left and right images as the raw signal for estimat- ing which side contained the target on a given trial. If the total luminance of the left image was higher than FIG. 3: The effect of lateral interactions on exam- ple images. Left column: black and white amoeba-target image (K = 4). Right column: Gray-scale natural image (the standard computer vision test image "Lena") after ap- plying a hard Difference of Gaussians (DoG) filter to enhance edges. Top row: Raw retinal input. Second row: Responses of orientation-selective elements before any lateral interactions (k = 0). To aid visualization, the activity of the maximally responding orientation-selective element at each pixel loca- tion is depicted as a gray-scale intensity. Rows 3-6: Activity after k = 1, 2, 3, 4 iterations of the multiplicative ODD kernel, as labeled. For each iteration, activity was multiplied by the local support, computed via linear convolution of the previ- ous output activity with the ODD kernel. Lateral interac- tions tended to support smooth contours, particularly those arising from amoeba segments, while suppressing clutter or background detail. 7 that of the right (relative to threshold), the response of the model would be reported as target on the left. Ide- ally, after several iterations of the ODD kernel, no seg- ments would remain in the distractor image and only amoeba segments would remain in the target image; in practice, the total luminance served as a measure of con- fidence. Given the much larger number of trials (1000) available for assessing model performance, 100 equally spaced threshold values were used to calculate the cor- responding ROC curves. As with the ROC curves con- structed from the confidence values reported by the hu- man subjects, the ROC curves computed from the confi- dence values reported by the model give the cumulative proportional true positive rate as a function of cumula- tive proportional false positive rate, with the confidence threshold varied from zero to maximum. Graphically, the area under the ROC curves is given by the amount of overlap between the total luminance histograms (see figure 4) for the target and distractor images [44]. ROC curves for human subjects show performance in- creasingly above chance, indicated by a diagonal line of slope 1, as a function of both increasing SOA and de- creasing target complexity. For amoeba targets of low to moderate complexity, ROC curves obtained from hu- man subjects were well matched to those generated by the model cortical association fields, consistent with the hypothesis that lateral interactions between orientation- selective neurons contribute to human contour percep- tion, at least for simple targets. The area under the ROC curve (AUC) gives the prob- ability that a randomly chosen target image will be cor- rectly classified relative to a randomly chosen distractor image, and thus provides a threshold-independent assess- ment of performance on the 2AFC task. Both the aver- age over human subjects and the model cortical associa- tion fields exhibited qualitatively similar performance on the 2AFC amoeba/no-amoeba task (Figure 7). Perfor- mance declined as a function of increasing target com- plexity, both for human subjects, measured at a fixed SOA, and for the model, measured at a fixed number of iterations, implying that K was an effective control parameter for adjusting task difficulty. At 20 ms SOA, the performance of human subjects was indistinguishable from chance, suggesting that our optimized masks effec- tively prevented the development of bottom-up cortical responses, even for the simplest targets (K = 2). Al- though some studies report that line drawings are pro- cessed more rapidly than natural images, with above chance performance being observed at short SOA val- ues [1, 26], the fact that performance on the amoeba/no- amoeba task was no better than chance at a 20 ms SOA implies that our optimized masks effectively interrupted visual processing of the amoeba targets. Since the model used here did not include any account for the time course of bottom-up retinocortical dynamics, we assumed that the performance of human subjects at 20 ms SOA should be equated to model performance at 0 iterations (prior to any lateral interactions), a time frame consistent with the FIG. 4: Histograms of total luminance in target and distractor images as a function of the number of it- erations. Red bins: Total activity histograms for all 1, 000 test target images. Blue bins: Total activity histograms for all 1, 000 test distractor images. The degree that the two distributions overlap is shown as the gray shaded area, which provides a measure of whether total luminance can be used to distinguish targets from distractors. The percentage in each shaded area shows the approximate lower bound amount of overlap of the two histograms, for comparison. Top row: To- tal summed activity over all retinal pixels. Little, if any bias between target and distractor images was evident in the in- put black and white images as there is nearly complete over- lap between the distributions. Subsequent rows: Total activ- ity histograms summed over all orientation-selective elements. Second row: Bottom-up responses prior to any lateral inter- actions. Third - sixth rows: Total activity histograms after 1 - 4 iterations of the multiplicative ODD kernel, respectively. Total summed activity became progressively more separable with additional iterations, as evinced by a decrease in the overlapping areas. 8 dial frequencies are processed by different cortical chan- nels [46]. Model performance might have been improved by training a new set of ODD kernels specifically for tar- gets containing K ≥ 6 radial frequencies, thereby utiliz- ing a hypothetical sub-population of orientation-selective neurons optimized for detecting high-curvature contours. Here, our model was limited to a single multiplicative kernel for detecting all predominately smooth contours. To quantify how average human performance on the 2AFC amoeba/no-amoeba task varied with SOA, and to compare with the dependence of model performance on the number of iterations of the ODD kernel, areas un- der both sets of ROC curves were fit to a monotonically increasing function of the following sigmoidal form: f (t) = F∞ 1 − (1 − 2F∞)e−λ(t−t0) . (1) For human experiments, the parameter t corresponds to the SOA in ms. Since we expect humans to perform close to 100% accuracy for very long SOA, we set F∞ = 1. Since humans perform essentially at chance (50%) for 20 ms SOA, we set t0 = 20 ms. Thus λ was the only free parameter; fits to the average human data were denoted by λH ; λH has units of 1/ms. Likewise, model perfor- mance was fit to a curve with the same functional form, with t = k measuring the number of iterations; λM was used to denote curve fits to the model data. However, visual inspection of the model data suggests that its per- formance saturates at less than 100% accuracy even after an infinite number of iterations, thus we forced the sig- moidal curve fit to the model results to asymptote at the final measured value of AUC: F∞ = AU Ck=4(K). Since the model performs better than chance after only 1 iter- ation, we set t0 = 0. For both the human experiments and the model performance, the functional form of f (t) ensures that f (t0) = 1 2 , corresponding to a minimal per- formance equal to chance. We find that 1/λH and 1/λM behave quite differently as a function of K, the number of radial frequencies used in amoeba generation (Figure 7). As anticipated for a re- laxational process governed by a single kernel, the model data was well described by a single value of λM (in units of 1/iterations), equal to 1.26. For the human subjects data, values of λH increased from 0.034 to 0.011 as a function of amoeba complexity, corresponding to lateral processing times of 29.8 to 90.6 ms, respectively. If hu- man performance depended on only a single set of lateral connections, then, at least in the linear approximation case, we might expect human performance to be well de- scribed by a single dominant time constant, represent- ing the dominant eigenmode of the horizontal interac- tions [47, 48]. Multiple time scales in the human perfor- mance case may emerge from any number of physiological mechanisms not included in the present model, including additional non-linearities in the action of the horizontal connections and/or contributions to contour perception from extrastriate areas. Our data do not allow us to make a firm distinction between these possibilities. FIG. 5: Psychophysical experiment schematic. The stimulus consisted of one target image and one distractor im- age (randomly positioned with equal probability on the left or right), presented simultaneously for an SOA between 20 ms and 200 ms, followed by an optimized 100 ms mask gen- erated from randomly rotated groups of target and distractor segments. Subjects indicated which side contained the target object (amoeba) using a computer mouse to click along a hor- izontal slider bar. Clicking far to the left or right indicated strong confidence that the corresponding side contained the target; clicking close to the center indicated weak confidence. A narrow gap in the center forced subjects to choose between left and right. distribution of the shortest measured response latencies recorded in primary visual cortex [45]. Overall, average human performance improved as a function of increasing SOA in a manner analogous to the improvement in model performance as a function of the number of iterations of the ODD kernel. This cor- respondence was especially evident for amoebas of low to moderate complexity (K ≤ 4 ). For more complex targets, model performance lagged well behind that of human subjects. Studies suggest that low and high ra- 9 FIG. 6: ROC curves comparing human and model performance on the amoeba/no amoeba task. Top two rows: ROC curves averaged over four different human test subjects using reported confidence (points). The dashed diagonal line in each plot indicates the curve corresponding to chance. Red, blue, green, black correspond to K = 2, 4, 6, 8, respectively. Bottom two rows: ROC curves for model cortical association fields computed from total activity histograms. However, one possible interpretation of the present re- sults is that the perception of simple contours is domi- nated by relatively fast lateral interactions placed early in the visual processing pathway, thereby accounting for the good fit between the model and experimental results for targets of low to moderate complexity. Building on this interpretation, we postulate that the perception of more complex contours requires more extensive, and therefore slower, processing mechanisms involving higher cortical areas, thus explaining the discrepancy between model and experimental performance as target complexity in- creases. Under the assumption that human perception of simple amoeba targets (K ≤ 4) depends primarily on re- current lateral interactions between orientation-selective neurons, we can estimate the time required for each it- eration of the multiplicative ODD kernel. This rate is estimated using the K = 2 time constants from the fits: λM,K=2/λH = 37.5 ms per iteration, a value consistent with estimates of lateral conduction delays within the same cortical area [13]. Having shown that the lateral interactions based on multiplicative ODD kernels can account for both spa- tial and temporal aspects human contour perception, we seek to identify model details that are essential to the performance reported here. First, we demonstrate that the proposed model is robust and does not require that the magnitude of the ODD kernel be carefully titrated to a precise value. Model performance on the 2AFC amoeba/no-amoeba task, measured by the area under the ROC curve (AUC) for increasing numbers of itera- tions (k = 0, 1, 2, 3, 4), was plotted for different values of the strength of the ODD kernel, given by the total in- tegrated strengths of the equal and opposite target and distractor contributions (Figure 8). The number of radial frequencies was fixed at K = 4. Qualitatively similar per- formance was obtained for ODD kernel strengths ranging from 300 to 400. The ODD kernel used in the present study, whose strength was set to 325, produced near opti- mal performance and also exhibited monotonic improve- ment with increasing numbers of iterations. That perfor- mance was generally insensitive to the value of the main free parameter in the model provides strong evidence for the robustness of the proposed contour detection mecha- nism based on multiplicative lateral interactions. A second aspect of the model that merits scrutiny is the detailed structure of the ODD kernels, which SOA=20msSOA=40msSOA=80ms01Falsepositiverate01TruepositiverateK=2K=4K=6K=8SOA=120msHumanPerformanceSOA=160msSOA=200msk=0k=1k=201Falsepositiverate01TruepositiverateModelPerformanceK=2K=4K=6K=8k=3k=4 10 FIG. 7: A comparison of human and model performance on the 2AFC amoeba/no amoeba task. Left: Average human performance for different SOA in milliseconds. Right: Performance of model cortical association fields for increasing numbers of iterations. Both panels: Accuracy, which is equivalent to area under the ROC curve, (error bars) fitted to single sigmoidal functions (solid lines). The four curves from top to bottom correspond to K = 2, 4, 6, 8 radial frequencies. were trained using computer-generated images in which the pairwise edge statistics uniquely identifying globally salient contours could be calculated directly. Previous models of contour perception typically employed much simpler patterns of lateral connectivity, in which excita- tory interactions were either collinear or cocircular, and inhibitory interactions were approximately independent of relative orientation [8, 24, 27, 47 -- 49]. To determine if the detailed structure of the ODD kernel was critical to the observed performance, we repeated the amoeba/no- amoeba experiment using a much simpler kernel whose basic form was consistent with a number of previously published models (see Figure 8). Specifically, we used a "Bowtie" kernel in which excitatory connections fanned out with an opening angle of π/6 and the difference in the preferred orientations of the pre- and post-synaptic elements differed by no more than ±π/6. Both exci- tatory and inhibitory connection strengths fell off in a Gaussian manner, with inhibition strength being insen- sitive to orientation. Although the overall accuracy of the Bowtie kernels was lower than that achieved by the ODD kernels, performance on the amoeba/no-amoeba tasks was qualitatively similar, particularly regarding the general monotonic improvement with the number of it- erations and the absence of a sensitive dependence on kernel strength. Thus, we conclude that multiplicative lateral interactions are able to preserve smooth closed contours while suppressing clutter in a manner that is robust to broad changes in model details. Discussion We have shown that simple models of neural activity in primary visual cortex, enriched with lateral associa- tion kernels, reproduce some of the behavioral features regarding the human perception of broken closed con- tours. Our results agree not only with the measured de- pendence on contour complexity but also with the tem- poral dependence of human perception as a function of SOA, suggesting that horizontal connections in V1 may play a non-trivial and global computational role in the perception of closed contours on very fast timescales. A number of studies relate to the potential contribu- tion of cortical association fields to human contour per- ception; these encompass a range of anatomical, physi- ological, psychophysical, and theoretical techniques [2 -- 5, 7 -- 10, 10, 11, 13, 16 -- 19, 50]. In particular, a number of theoretical models have sought to account for human contour perception at the level of biologically-plausible neural circuits [8, 27, 28, 49, 51 -- 54], with most studies incorporating some form of cortical association field con- figured to reinforce smoothness [24]. Although biologi- cally plausible models of cortical association fields have been used to account for the dependence of contour vis- ibility on key parameters controlling task difficulty, such as smoothness, closure, and density of background clut- ter [8], model cortical association fields have not been directly compared to the time course of human contour perception as a function of contour complexity. Here, we used cortical association fields based on ODD ker- nels, which were computed from differences in the pair- wise coactivation statistics of orientation-selective ele- ments arising from target as opposed to distractor im- ages. While we designed the kernels specifically for the amoeba-clutter disambiguation, we emphasize that the algorithm for the ODD kernel construction is completely general and can be used to improve detection of salient image features in any situation where generative mod- els of targets and distractors are known, or there exists data sets of sufficient size to characterize the contour co- occurrence statistics empirically for both targets and dis- tractors. In our experiments, ODD kernels were able to account for the experimentally observed variations in the saliency of closed contours as a function of parametric complexity and for the time course with which smooth contours are processed by cortical circuits. Crucial for these results was our use of a synthetic target/distractor data set with controllable complexity and the absence of top-down contextual features or local cues that might 04080120160200SOA(ms)0.500.751.00AccuracyK=2K=4K=6K=8Human01234IterationsK=2K=4K=6K=8Model give away target presence. Here, we used a semi-supervised training scheme to learn lateral connectivity patterns optimized for perform- ing the amoeba/no-amoeba task. Necessarily, we sought to model only a subset of the lateral interactions be- tween orientation-selective neurons, namely, those hori- zontal connections configured to reinforce smooth, closed contours. We did not attempt to capture the full range of spatial relationships between features extracted at early cortical processing stages [24, 55]. Presently, databases containing sufficient numbers of fully annotated and seg- mented natural images needed to reproduce the weeks (or months) of visual experience required to train the full complement of horizontal connections in the primary visual cortex do not exist. Moreover, the computational resources to exploit such databases, even if they did ex- ist, are highly non-trivial to assemble. Thus, we focused here on a subset of horizontal connections for which it was possible to construct synthetic surrogate images. At most, the proposed model represents a subset -- and only a subset -- of the lateral connections between orientation- selective cortical neurons. Moreover, even a complete set of such horizontal connections would, at most, represent but a subset of the cortical mechanisms that contribute to the time course and shape-dependence of contour per- ception. FIG. 8: A comparison of ODD and simpler "Bowtie" kernel performance on the on the 2AFC, K = 4 amoeba/no amoeba task plotted as a function of the number of iterations for a range of different kernel strengths. Line width and marker size denote values on kernel strength, which was the main free parameter in the model. Kernel strength is a dimensionless constant. Black lines: ODD kernel performance. Blue lines: "Bowtie" ker- nel performance. Qualitative behavior was similar for both kernels, demonstrating that multiplicative lateral interactions act robustly to reinforce smooth closed contours. The supervised training scheme employed here might be related to perceptual learning phenomena, which take place over time scales much shorter than those typi- cally associated with developmental processes [56 -- 58]. It is possible that known physiological mechanisms, such 11 as spike-timing-dependent plasticity (STDP), especially with accounts for realistic conduction delays [59], could mediate a rapid refinement of lateral connections so as to facilitate the perception of amoeba targets. Moreover, physiological plasticity mechanisms might produce differ- ent patterns of connectivity for orientation-selective ele- ments representing points of low as opposed to high local curvature, thereby optimizing lateral interactions for con- tours of varying complexity. Here, we made no attempt to customize distinct ODD kernels for detecting contours of varying complexity. Instead, a single ODD kernel was trained using a complete set of images in which differ- ent numbers of radial frequency components were equally represented. Although we did not investigate whether, or to what extent, the performance of human subjects im- proved over the course of the amoeba/no-amoeba experi- ment, such investigations might shed insight into the role of perceptual learning in the detection of closed contours. The question of how lateral connectivity based on ODD kernels might be acquired during development was not addressed explicitly. In principle, coactivation statistics between pairs of orientation-selective neurons could be accumulated over time in an unsupervised manner by a Hebbian-like learning rule [60]. Under natural view- ing conditions, we expect that contour fragments consis- tent with smooth, closed boundaries would tend to occur simultaneously, whereas contour fragments inconsistent with object boundaries would tend occur at random tem- poral delays. Thus, a Hebbian-like learning rule sensitive to temporal correlations, such as certain mathematical forms of STDP-like learning rules [61], might under nor- mal developmental conditions lead to connectivity pat- terns that reinforce smooth contours. Of course, human contour perception may have noth- ing to do with cortical association fields, or lateral in- teractions may play a subordinate role. Early mod- els showed how spatial filtering could enhance texture- defined contours in the absence of orientation-specific interactions [4] and short-range lateral interactions can accentuate texture-defined boundaries [31, 62]. How- ever, psychophysical studies employing implicit con- tours [2, 7, 8], in which foreground and background ele- ments are present at equal density and which lack explicit texture cues, appear to rule out explanations that omit long-range, orientation-specific interactions. An influen- tial class of biologically-inspired computer vision models achieves a degree of viewpoint-invariant object recogni- tion by constructing feed-forward hierarchies to extract progressively more complex and viewpoint invariant fea- tures [33, 63]. By analogy with such models, scale- and position-independent representations for detecting long, smooth contours could in principle be constructed hier- archically, starting with simple edge detectors and build- ing up progressively longer, more complex curves using a "bag-of-features" approach. Presently, there appear to be insufficient data to decide whether human con- tour perception involves primarily lateral, feed-forward, or even top-down connections [30, 32, 64]. Hypothet- 01234Iterations0.500.751.00AccuracyODDBowtie30032540075125175 ically, the cortical association fields used in the present study could have been implemented as a feed-forward ar- chitecture, using a hierarchy of orientation-selective neu- rons to link progressively more widely separated contour fragments. Functionally, there may not exist a clean distinction between lateral, feed-forward and feed-back topologies, with the possibility that all three types of connectivity contribute to human contour perception. To quantify the temporal dynamics underlying visual processing, we performed speed-of-sight psychophysical experiments that required subjects to detect closed con- tours (amoebas) spanning a range of shapes, sizes and positions, whose smoothness could be adjusted paramet- rically by varying the number of radial frequencies (with randomly chosen amplitudes). To better approximate natural viewing conditions, in which target objects usu- ally appear against noisy backgrounds and both fore- ground and background objects consist of similar low- level visual features, our amoeba/no-amoeba task re- quired amoeba targets to be distinguished from locally indistinguishable open contour fragments (clutter). For amoeba targets consisting of only a few radial frequen- cies (K ≤ 4), human subjects were able to perform at close to 100% accuracy after seeing target/distractor im- age pairs for less than 200 ms, consistent with a number of studies showing that the recognition of unambiguous targets typically requires 150 − 250 ms to reach asymp- totic performance [22, 23, 35], here likely aided by the high intrinsic saliency of closed shapes relative to open shapes [7]. Because mean inter-saccade intervals are also in the range of 250 ms [34], speed-of-sight studies indi- cate that unambiguous targets in most natural images can be recognized in a single glance. Similarly, we found that closed contours of low to moderate complexity read- ily "pop out" against background clutter, implying that such radial frequency patterns are processed in parallel, presumably by intrinsic cortical circuitry optimized for automatically extracting smooth, closed contours. As saccadic eye movements were unlikely to play a signif- icant role for such brief presentations, it is unclear to what extent attentional mechanisms are relevant to the speed-of-sight amoeba/no-amoeba task. Our results further indicate that subjects perform no better than chance at SOAs shorter than approximately 20 ms. Other studies, however, report above chance performance on unambiguous target detection tasks at similarly short SOA values [1, 23, 26, 33]. The discrep- ancy may be attributed to the different masks employed. Whereas the above cited studies used masks consisting of either spatially filtered (e.g. 1/f ) noise, distractor images, or scrambled versions of the target image set, we constructed rotation masks that were optimized for each target/distractor image pair [20]. Our working hy- pothesis was that an optimized mask should completely obscure the target object in the target-mask composite image; also referred to as pattern masking. The require- ment that the mask completely hide the target follows from the assumption that at very short SOA, the tar- 12 get and mask images are likely to be effectively fused due to the finite response time of neurons and recep- tors in the early visual system [65]. For the amoeba/no- amoeba task, we created optimized masks by rotating the amoeba and clutter fragments with the goal of produc- ing the maximum amount of interference in the responses of orientation-selective cells. Presumably, maximum in- terference occurs when orientation-selective neurons are presented with randomly rotated contour fragments in rapid succession. Although backward masks can have heterogeneous effects, with performance in some cases showing a U -shaped dependence on SOA [66], for the masks used here performance always increased monoton- ically with SOA. Empirically, the fact that performance was no better than chance at 20 ms SOA suggests that our optimized masks were able to effectively interrupt the processing of smooth, closed contours at early corti- cal processing stages. Indeed, the ability to drive overall performance down to chance at SOA values shorter than 20 ms could provide an operational criteria for assessing the degree to which a given backward pattern mask is able to effectively interrupt visual processing. The amoeba/no-amoeba task required the integration of information over length scales spanning viewing angles of approximately 3 − 4◦, larger than the classical excita- tory receptive field size of parafoveal V1 neurons. The amoeba/no-amoeba image set (see Figure 1) was con- figured so that purely local information, such as a few adjoining contour fragments, would not be sufficient to solve the target detection problem. Rather, distinguish- ing amoebas from clutter required integrating global in- formation across multiple contour fragments. Our re- sults suggest that such global integration can be accom- plished via lateral interactions between local, orientation- selective filters. Although the density of target and clutter segments was not precisely equilibrated in our amoeba/no-amoeba image set, the wide range of target sizes and shapes spanned by our image generation algo- rithm makes it unlikely that the near perfect performance of human subjects at long SOA could have been attained using density cues alone [4]. Here, lateral inputs were used to modulate the bottom-up responses in a multi- plicative fashion, so that our cortical association fields acted primarily as gates that suppressed contour frag- ments that did not receive sufficiently strong contextual support. By preventing lateral inputs from producing ac- tivity unless there was already a strong bottom-up input, a multiplicative non-linearity prevented the activation of contour fragments not present in the original image. The phenomenon of illusory contours suggests that in some cases contextual effects can produce activity even in the absence of a direct bottom-up response [30]. The pre- cise form of the multiplicative interaction used here was adopted for algorithmic simplicity rather than for biolog- ical realism. We observed that including a small additive contribution from the lateral interactions did not funda- mentally affect our conclusions. This suggests that ODD kernels, if implemented more generally, might account for the perception of illusory contours as well. However, a more realistic description of the underlying cellular and synaptic dynamics would likely be necessary to model a relaxation process that includes both additive and mul- tiplicative elements. Both the model and the psychophysical experiments employed a 2AFC design (see Figure 5) in which the goal was to correctly identify which of a pair of images contained an amoeba target. Since each trial involved a forced choice between two images, the model used a simple classifier that labeled the image with greater to- tal activity as the target. For both human subjects and the model, the number of radial frequencies K proved to be a good control parameter for adjusting task difficulty (see Figure 7). For targets of low to moderate complex- ity, both model performance (as a function of number of iterations) and human performance (as a function of increasing SOA) monotonically approached nearly per- fect asymptotic performance as described by a single sig- moidal function with a characteristic scale, representing either time or number of iterations, that increased with K (see Figure 7). Based on comparison with human perfor- mance at different SOA values, each iteration of the ODD kernels was estimated to require approximately 37.6 ms of cortical processing time, consistent with measured con- duction delays between laterally connected cortical neu- rons [13]. Prior to any lateral interactions, the stimulus was pro- jected onto a retinotopic array of orientation-selective fil- ter elements, providing a convenient representation for learning cortical association fields by computing differ- ences in pairwise coactivation statistics between target and distractor images. We found that each iteration of the ODD kernel increased the activity of contour frag- ments that were part of amoebas compared to the activ- ity of clutter fragments, so that after several iterations the mean overall activity, summed across all orientation- selective filter elements, was higher on average for target images than distractor images (see Figure 4). Even in tri- als that were incorrectly classified, contour fragments be- longing to amoebas were typically still favored relative to background clutter. Because the total number of contour fragments varied from trial to trial, with only the average number of fragments being fixed across the entire image set, our relatively crude criterion for discriminating be- tween target and distractor images sometimes led to clas- sification errors even when amoeba fragments had been partially segmented from the background clutter, sim- ply because the distractor image initially contained more fragments. A more sophisticated classifier might have led to a closer correspondence between model and hu- man performance. Although performance of the present multiplicative model appeared to saturate after only a few iterations of the ODD kernel (e.g. k ≤ 4), it is possi- ble that a different implementation might have continued to show improvements after additional iterations. How- ever, the longer processing time implied by additional iterations suggests that other physiological mechanisms, 13 particularly visual search, would likely come into play. Granted, there is an apparent mismatch between the fading of clutter elements in the model and the persis- tence of such elements perceptually in human subjects. To reconcile this apparent mismatch, it has been sug- gested that the initial perception of brightness might be driven by the initial bottom-up response of the individual orientation-selective feature detectors, whereas persistent responses across these same feature detectors might drive salience [28]. The amoeba/no-amoeba image set was designed to al- low for parameterized complexity (in terms of the amount of clutter, number of radial frequencies, etc.) while avoid- ing reference to exogenous world knowledge. Since the amoeba/no-amoeba image set was machine generated, it was possible to produce a very large number of train- ing images; 40, 000 target and 40, 000 distractor images at 256 × 256 pixel resolution were used to train ODD kernels in the present study. Many computer vision sys- tems employ standard image classification datasets such as the Caltech 101 [67], which allows for uniform bench- marking and thus facilitates direct comparison between models. Datasets based on natural images, however, suf- fer from several shortcomings. First, the resolution and number of images are fixed when the set is created. While some man-made datasets, such as MNIST [68]), consist of tens of thousands of handwritten characters, anno- tated sets of natural photographs ideal for speed-of-sight experiments are typically limited to a few hundred im- ages. In contrast, humans are exposed to millions of natural scenes during visual development. Biologically motivated models that attempt to replicate human per- formance might require similar numbers of examples. A second shortcoming of natural image datasets is preva- lence of high-level contextual information that utilizes ex- ogenous world knowledge, such as the increased a priori likelihood of finding a car on a road, or an animal in a for- est. Exploiting such exogenous world knowledge posses a formidable challenge for existing computational models and, on tasks that employ natural images, may obscure the ability of such models to extract behaviorally mean- ingful information from low-level visual cues. Third, nat- ural image datasets typically provide limited capability for adjusting intrinsic task difficultly. For example, one widely used dataset [33] includes photographs of animals at different distances, but only a few discrete distances are annotated and the relationship of target distance to task difficultly is not easily quantified. Here, we illus- trated how a synthetic set of images could be used to compare model and human performance in a task with parametric difficulty, potentially validating the use of ar- tificial as opposed to natural images. The present study addressed the role of cortical associ- ation fields in the perception of closed contours, which are presumably important for detecting visual targets based on shape or outline. Although studies show that human subjects can rapidly distinguish between images contain- ing target and non-target object categories using only the line drawings obtained by filtering natural scenes [1], normal experience involves a number of complementary visual cues, such as texture, color, motion and stereop- sis. Presumably, cortical association fields also act to re- inforce features representing these complementary visual cues as well. Human subjects, for example, can distin- guish whether pairs of texture patches were drawn from the same natural object or two different natural objects in a manner that exhibits a similar dependence on pair- wise co-occurrence statistics as was found for orientated edges [55]. We may speculate that an analysis of coacti- vation statistics for features selective to a combination of cues such as local orientation, texture, color, motion, and disparity may lead to a more general and more powerful set of kernels capable of fast and effective determination of global object properties, which in turn can play an important role in complex object identification. Methods Synthetic amoeba/no-amoeba image set An amoeba is a type of radial frequency pattern [36] consisting of a deformed circle in which the radius varies as a function of the polar angle. By choosing the number and relative amplitudes of the different frequency com- ponents, the radius can describe an arbitrarily complex shape, exactly analogous to how a Fourier basis can be used to construct an arbitrary waveform on a finite in- terval. Each radial frequency component was represented by a sinusoidal function defined at C = 1024 discrete po- lar angles, spaced uniformly on the interval [0, 2π). The cutoff radial frequency used in constructing the closed contour provided a control parameter for regulating the complexity of the resulting figure, which ranged from nearly circular, when only the 2 lowest radial frequen- cies had non-zero amplitudes, to highly sinusoidal and irregular, when the first 8 radial frequencies had non- zero amplitudes. All amoeba shapes generated here may be considered smooth, in that local curvature was always bounded. In detail, the radius of an amoeba at each polar angle was: r(φc) = A0 + K(cid:88) n=1 (cid:18) 2πn C (cid:19) An cos + αn . (2) All amplitudes An were initially drawn from normal dis- tributions with 0 mean and unit variance. All phases αn were drawn from uniform distributions over the in- terval 0 and π/2. The resulting radial frequency pattern was then linearly rescaled so that the maximum radius, rmax, was equal to a random number drawn from a uni- form distribution such that L/4 ≤ rmax ≤ L/2, where L is the linear size of the square image (L = 256 pixels), and the minimum radius was given by a second randomly chosen value so that rmax/4 ≤ rmin ≤ rmax/2. Uniform 14 pseudo-random numbers were generated by the intrinsic MATLAB 7.0 function RAND, or its Octave 3.2.3 equiv- alent. To facilitate the construction of locally indistinguish- able clutter and model contour occlusion in natural im- ages, amoeba contours were divided into 16 periodically- spaced fragments by removing short sections whose lengths varied within a specified range. Specifically, the gaps between amoeba fragments varied from 16 to 32 in units of discrete polar angle (2π/C). Amoeba contours were then broken into fragments by periodically inserting 16 gaps of variable width ranging from 16 to 32, spaced C/16 segments apart. Gaps were deleted from the un- derlying contour, so that the polar angle subtended by each fragment varied in accordance with the changes in preceding gap width. The starting point of the first gap was chosen randomly on the interval 1, C/G, so that over the entire image set the inserted gaps were distributed uniformly around the circle. To create clutter fragments, an amoeba was first gen- erated using the above procedure. Consecutive amoeba fragments were then grouped, with the number of frag- ments in each group determined by a Poisson process with a mean value of 2 and an upper cutoff of 3. Each group of amoeba fragments was then rotated about its center of mass through random angles on the interval π/8 to 7π/8. The resulting clutter consisted of the same fragments as the original amoeba but rotated so that collectively the rotated fragments no longer supported the perception of a closed object. Clutter fragments constructed in this manner were thus locally indistin- guishable from amoeba fragments. To create clutter in both target and distractor images, several amoebas were first superimposed at random positions and then groups of fragments rotated following the procedure described above. All amoebas contained the same total number of contour fragments (and therefore the same number of gaps) but varied in both maximum diameter and total contour length. The center of each amoeba was chosen randomly un- der the restriction that no contour be allowed to cross an image boundary. Specifically, the x-coordinate of the amoeba center, x0, was chosen randomly on a restricted interval, rmax ≤ x0 ≤ L − rmax, and likewise for the y- coordinate, y0. When groups of amoeba fragments were randomly rotated to make clutter, portions of a contour belonging to a clutter fragment would occasionally cross an image boundary. In such cases, any out-of-bounds portions of a contour were reflected back into the image region using mirror boundary conditions. Target images always consisted of 1 set of amoeba frag- ments and 2 sets of clutter fragments. Distractor images consisted of 3 sets of clutter fragments and thus, averaged over the entire image set, had the same mean luminance and the same variance as the target images. Mask images were constructed following a procedure nearly identical to that used for constructing distractor images, except that mask images consisted of 6 sets of clutter fragments, obtained by randomly rotating the 6 original amoeba ob- jects used in constructing the corresponding target and distractor images. All contour fragments were initially represented as a set of points in polar coordinates, cor- responding to the radius at each discrete polar angle. Points along the contour were then transformed back to Cartesian coordinates and rounded to the nearest dis- crete pixel value. MATLAB scripts for generating the image set used in this study are publicly available at: http://petavision.sourceforge.net. Ethics statement The Los Alamos National Laboratory (LANL) Human Subjects Research Review Board (HSRRB) has reviewed the following experimental protocol and determined that it provides adequate safeguards for protecting the rights and welfare of human subjects involved in the protocol. The protocol was reviewed and approved in compliance with the U.S. Department of Health and Human Ser- vices (DHHS) regulations for the Protection of Human Subjects, 45 CFR 46, and in accordance with the LANL Federal Wide Assurance (FWA#00000362) with the Na- tional Institutes of Health/Office for Human Research Protections (NIH/OHRP). The identification number is LANL 08-03 X. Human psychophysics evaluated using Human performance was two- alternative forced choice (2AFC) psychophysical exper- iments. There were 5 subjects, all with normal or corrected-to-normal vision. One subject only contributed data for a portion of the tested SOAs. Each subject was seated in a dark room, at an approximate distance of 65 cm from a 19-inch nominal (36.2× 27 cm actual size) Hi- tachi 751 CRT monitor. Images spanned a viewing angle of approximately 7◦ × 7◦. The monitor resolution was 1024 × 768 pixels and the refresh rate was 100 Hz. The display was driven by a dual-core 3.0 GHz Mac Pro, with MATLAB 7.6 running Psychtoolbox [69]. After a short training period to familiarize the subject with the task, one target image and one distractor image were shown side by side, followed by a mask intended to interrupt cognitive processing of the target and distrac- tor images. Two separate sets of experiments were con- ducted for each subject. In one set, the SOA was chosen randomly from the values 40, 80, 120 ms. For the second set of experiments, the SOA was chosen randomly from the values 20, 160, 200 ms. The duration of the stimulus was always the same as the SOA, and thus both the target and distractor images remained visible until mask onset. The duration of the mask was always 100 ms. Each sub- ject was shown 1200 images divided into 10 blocks of 120 images, with rest breaks in between blocks (rest break du- ration was at the discretion of each subject). The pace of 15 the experiment was under the control of the subject, who initiated each trial using the space bar. A small temporal jitter, chosen uniformly between 0 to 250 ms, was added to the interval preceding each trial, to prevent entrain- ment. Task conditions, consisting of variations in both the SOA and the number of radial frequencies K, were randomly interleaved such that each condition occurred the same number of times over the course of the entire experiment. On each trial, subjects indicated which side contained the target, using a mouse-driven slider bar to report con- fidence (see Figure 5). The reported confidence values were used to construct receiver operating characteristic (ROC) curves, which plot the percentage of true posi- tives (or hits) against the percentage of false positives (or false alarms), with each true/false positive pair obtained by setting a confidence threshold at a different location along the slider bar. A correct response was not neces- sarily considered a true positive: to generate one point on the ROC curve, the reported confidence on each trial was measured relative to the current threshold position, which could be to either the left or to the right of center. Thus, a trial might be labeled as incorrect, even though the subject moved the slider bar in the correct direction, as long as the threshold level was not exceeded. Specifi- cally, whenever the reported confidence fell to the left of threshold, the corresponding trial was treated as though the subject reported the target as being to the left, even if the threshold location had been set to the right of cen- ter and the confidence bar had actually been slid to the right. Likewise, when the reported confidence fell to the right of the current threshold position, the trial was al- ways treated as if the subject had reported the target to the right, again regardless of how the subject moved the slider bar relative to the center position. By choosing a range of threshold positions, spanning the full range of reported confidence values, a complete ROC curve was obtained. Note that as the threshold was moved closer to the left edge of the slider bar, the percentage of true and false positives both approached minimum val- ues, since only trials with very high reported confidence could contribute to either the true positive or false posi- tive rate (most trials were rejected as either true or false negatives). As the threshold position moved closer to the center of the confidence slider bar, the percentage of true positives increased. Finally, as the threshold was moved closer to the right edge of the slider bar, both the true positive rate and the percentage of false posi- tives approached maximum values. The true positive rate averaged over all false positive rates, or the area under the ROC curve (AUC), was used as an overall measure of subject performance. The AUC is equivalent to the probability that a randomly chosen target image will be correctly classified relative to a randomly chosen distrac- tor image, and thus directly predicts performance on the 2AFC task. Results for each SOA and for each value of K were averaged over 5 subjects. Error bars denote the standard deviation over the 5 subjects. Model rotation matrix, Model cortical association fields were based on differ- ences in the coactivation statistics of orientation-selective filter elements drawn from target and distractor images. Geisler and Perry measured co-occurrence statistics for oriented edges in human segmented natural images [25], and found a close correspondence to human judgments as to whether pairs of short line fragments were drawn from the same or different contours. Thus, we refer to the difference in coactivation statistics between target object and distractor images as Object-Distractor Dif- ference (ODD) kernels. ODD kernels were trained using 40, 000 target and 40, 000 distractor images, each divided into 4 sets of 10, 000 images each, with each set associ- ated with a different value of K = 2, 4, 6, 8. The order in which the images were presented had no bearing on the final form of the ODD kernel; that is, there was no tem- poral component to the training. Training with more images did not substantively improve performance, al- though small differences were observed in the ODD ker- nels trained using a smaller number of images (10, 000 target and 10, 000 distractor images). Each 256×256 pixel training image activated a regular array of 256×256 retinal elements whose outputs were ei- ther 0 or 1, depending on whether the corresponding im- age pixel was ON or OFF, respectively. Each retinal unit activated a local neighborhood of orientation-selective fil- ters, which spanned 8 angles spaced uniformly between 0 and π. To mitigate aliasing effects, the orientation- selective filters were rotated by a small, fixed offset, equal to π/16, relative to the axis of the training images. All orientation-selective filters were 7×7 pixels in extent and consisted of a central excitatory subunit, represented by an elliptical Gaussian with a standard deviation of 7.0 in the longest direction and an aspect ratio of 7.0, flanked by two inhibitory subunits whose shapes were identical to the central excitatory subunit but were offset by ±1.4 pixels in the direction orthogonal to the preferred axis. The weight Wθ(x1−x2, y1−y2), from a retinal element at (x2, y2) to a filter element at (x1, y1) with dominant orientation θ, was given by a sum over excitatory and inhibitory subunits: W θ(x1 − x2, y1 − y2) = W θ(r1 − r2) (cid:20) 1 (cid:0)r1 − r2 (cid:0)r1 + f − r2 (cid:0)r1 − f − r2 the matrix σ =(cid:2) 1 0 (cid:1)T(cid:21) (cid:1) · R−1 θ σ−1Rθ ·(cid:0)r1 − r2 (cid:1)T(cid:21) θ σ−1Rθ ·(cid:0)r1 + f − r2 (cid:1) · R−1 (cid:1)T(cid:21)(cid:41) (cid:1) · R−1 θ σ−1Rθ ·(cid:0)r1 − f − r2 (cid:3) describes the shape of the elliptical where the position vector is given by ri = [xi, yi] and (cid:20) 1 (cid:20) 1 − exp − exp (cid:40) (3) = A exp 2 2 2 Gaussian subunits for θ = 0. In Eq. 3, Rθ is a unitary 0 7 (cid:20) cos θ sin θ − sin θ cos θ (cid:21) , Rθ = 16 (4) and f = [0.0, 1.4] is a translation vector in the direc- tion orthogonal to the dominant orientation when θ = 0. The amplitude A was determined empirically so that the total integrated strength of all excitatory connections made by each retinal unit equaled 20.0 (and thus the total strength of all inhibitory connections made by each retinal unit equaled −40.0). Mirror boundary conditions were used to mitigate edge effects. The retinal input to each orientation-selective filter element s(x1, y1, θ) was then given by s(x1, y1, θ) = W θ(x1 − x2, y1 − y2)I (x1,y1)(x2, y2), (cid:88) x2,y2 (5) where I (x1,y1) is the 7 × 7 binary input image patch cen- tered on (x1, y1). The sum is over all pixels (x2, y2) that are part of this image patch. The initial output of each orientation-selective filter element z0(x1, y1, θ) was ob- tained by comparing the sum of its excitatory and in- hibitory retinal input to a fixed threshold of 0.5. Values below threshold were set to 0 whereas values above unity were set to 1.0. Thus is an element-wise implementation of these thresholds. The responses of all suprathreshold orientation-selective filters contributed to the coactivation statistics, with only the relative distance, direction, and orientation of filter pairs recorded. Because of the threshold condition, only the most active orientation-selective filters contributed to the coactivation statistics. For every suprathreshold filter element extracted from the i-th target image, coactivation statistics were accu- mulated relative to all surrounding suprathreshold filter elements extracted from the same image. Thus the ODD kernel G is given by (cid:0)ρ(x−x0, y−y0), φθ0(x−x0, y−y0)(cid:1) = (cid:88) z0(x, y, θ), Gti θ−θ0 x,y , (8) where the radial distance ρ is a function of the (x, y) coordinates of the two filter elements, the direction φ is the angle measured relative to θ0, the sum is over all suprathreshold elements within a cutoff radius of 32, the superscript ti denotes the i-th target image, and the dif- ference in the orientations of the two filter elements θ−θ0 is taken modulo π. Because the amoeba/no-amoeba im- age set was translationally invariant and isotropic, the where the function, z(x1, y1, θ) = g(cid:0)s(x1, y1, θ)(cid:1), 0 s < 0.5 0.5 ≤ s ≤ 1.0 s > 1.0 g(s) = s 1 , (6) (7) central filter element may without loss of generality be shifted and rotated to a canonical position and orienta- tion, so that the dependence on ρ0, φ0, θ0 may be omitted. The coactivation statistics for the i-th target image can then be written simply as Gti θ (ρ, φ), where (ρ, φ) gives the distance and direction from the origin to the filter el- ement with orientation θ, given that the filter element at the origin has orientation π/16. An analogous expression gives the coactivation statistics for the j-th distractor im- age Gdj θ (ρ, φ). The ODD kernel Gθ(ρ, φ) is given by the difference (cid:88) Gθ(ρ, φ) = W+ (cid:88) θ (ρ, φ) − W− Gti Gdj θ (ρ, φ), (9) i j where the sums are taken over all target and distractor images and the normalization factors W+ and W− are de- termined empirically so as to yield a total ODD strength of 325 (see Figure 8 and Results), defined as the sum over all ODD kernel elements arising from either the target or distractor components. By construction, the sum over all ODD kernel elements equals zero, so that the average lateral support for randomly distributed edge fragments would be neutral. Our results did not depend critically on the RMS magnitude of the ODD kernel (see Figure 8). To minimize storage requirements individual connection strengths were stored as unsigned 8-bit integers, so that the results of the present study did not depend on com- putation of high precision kernels. As described above, the canonical ODD kernel is de- fined relative to filter elements at the origin with orienta- tion π/16. Filter elements located away from the origin can be accounted for by a trivial translation. To account for filter elements with different orientations, separate ODD kernels were computed for all 8 orientations then rotated to a common orientation and averaged to pro- duce a canonical ODD kernel. The canonical kernel was then rotated in steps between 0 and π (offset by π/16) and then interpolated to Cartesian x−y axes by rounding to the nearest integer coordinates. Although it has been demonstrated that global contour saliency is enhanced for orientations along the cardinal axes [58], this bias is by construction absent from this model. ODD kernels were used to compute lateral support for 17 each orientation-selective filter element, via linear convo- lution. The output of each filter element was then modu- lated in a multiplicative fashion by the computed lateral support. The procedure was iterated by calculating new values for the lateral support s, which were again used to modulate filter outputs in a multiplicative fashion: Gθ(ρ, φθ)zk−1(x(cid:48), y(cid:48), θ(cid:48) sk(x, y, θ) = zk−1(x, y, θ) ), (cid:88) x(cid:48),y(cid:48),θ(cid:48) (10) where the subscript k denotes the k-th iteration. The same kernel was used for all iterations. All source code used to train and apply cortical association fields is pub- licly available at http://sourceforge.net/projects/petavision/. elements, T = (cid:80) To measure model performance, in each trial 1 target image and 1 distractor image were tested as a pair, so as to emulate the 2AFC format of the human experi- ments. The orientation-selective filter responses to both test images were evaluated after k = 0, 1, 2, 3, 4 iterations of the ODD kernel. The total activation across all filter x,y,θ(cid:48) zk(x, y, θ(cid:48)), was used to compare the two test images. Since the model cortical association fields tended to support contour fragments belonging to amoebas while inhibiting clutter fragments, the image with higher total activation T was assumed to be the target image. Error bars for the model performance (as shown in Figure 7) were estimated using the standard deviation of a binomial distribution with probability p equal to percent correct and N equal to the number of trials. Acknowledgments The authors wish to thank Steven Zucker for stimu- lating discussions that helped initiate this project. This work was supported by Los Alamos National Laboratory LDRD program under project 20090006DR; the National Science Foundation, grant ID 0749348; and the DARPA NeoVision2 project. This publication qualified for un- classified release under DUSA BIOSCI with LA-UR 11- 00499. [1] L. Velisavljevi´c and J. H. Elder, J Vision 9 (2009). [2] D. J. Field, A. Hayes, and R. F. Hess, Vision Res 33, 173 Computation. [7] I. Kov´acs and B. Julesz, P Natl Acad Sci USA 90, 7495 (1993), ISSN 0042-6989. (1993). [3] G. Loffler, Vision Res 48, 2106 (2008), ISSN 0042-6989, [8] M. W. Pettet, S. P. McKee, and N. M. Grzywacz, Vision vision Res Reviews. Res 38, 865 (1998), ISSN 0042-6989. [4] R. Hess and D. Field, Trends Cogn Sci 3, 480 (1999), [9] U. Polat and D. Sagi, Vision Res 33, 993 (1993). ISSN 1364-6613. [10] M. K. Kapadia, M. Ito, C. D. Gilbert, and G. West- [5] D. Fitzpatrick, Curr Opin in Neurobiol 10, 438 (2000), heimer, Neuron 15, 843 (1995). ISSN 0959-4388. [11] U. Polat, A. Terkin, and O. Yehezkel, Adv Cogn Psych [6] P. Seri´es, J. Lorenceau, and Y. Fr´egnac, J Physiology- Paris 97, 453 (2003), ISSN 0928-4257, neuroscience and 3, 153 (2008). [12] P.-C. Huang and R. F. Hess, Vision Res 47, 3108 (2007). 18 [13] V. Bringuier, F. Chavane, L. Glaeser, and Y. Fr´egnac, J. Neurosci. 26, 8254 (2006). Science 283, 695 (1999). [40] J. P. Jones and L. A. Palmer, J. Neurophys. 58, 1233 [14] J. R. Cavanaugh, W. Bair, and J. A. Movshon, J. Neu- (1987). rophys. 88, 2530 (2002). [41] T. W. Troyer, A. E. Krukowski, N. J. Priebe, and K. D. [15] J. R. Cavanaugh, W. Bair, and J. A. Movshon, J. Neu- Miller, J Neurosci 18, 5908 (1998). rophys. 88, 2547 (2002). [42] A. Angelucci, J. B. Levitt, E. J. S. Walton, J.-M. Hup´e, [16] A. Pooresmaeili, J. L. Herrero, M. W. Self, P. R. Roelf- J. Bullier, and J. S. Lund, J Neurosci 22, 8633 (2002). sema, and A. Thiele, J. Neurosci. 30, 12745 (2010). [43] P. Azzopardi and A. Cowey, P Natl Acad Sci USA 94, [17] W. H. Bosking, Y. Zhang, B. Schofield, and D. Fitz- 14190 (1997). patrick, J. Neurosci. 17, 2112 (1997). [18] C. Gilbert and T. Wiesel, J. Neurosci. 9, 2432 (1989). [19] R. Malach, Y. Amir, M. Harel, and A. Grinvald, P Natl [44] N. A. Macmillan and C. D. Creelman, Detection theory: a user's guide (CUP Archive, Cambridge, 1991). [45] J. H. R. Maunsell and J. R. Gibson, J. Neurophys. 68, Acad Sci USA 90, 10469 (1993). 1332 (1992). [20] R. F. Hess, W. H. A. Beaudot, and K. T. Mullen, Vision [46] J. Bell, D. R. Badcock, H. Wilson, and F. Wilkinson, Res 41, 1023 (2001), ISSN 0042-6989. [21] C. Keysers, D.-K. Xiao, P. Foldi`ak, and D. I. Perrett, J Cognitive Neurosci 13, 90 (2001). [22] C. Keysers and D. I. Perrett, Trends Cogn Sci 6, 120 (2002), ISSN 1364-6613. [23] N. Bacon-Mac´e, M. J.-M. Mac´e, M. Fabre-Thorpe, and S. J. Thorpe, Vision Res 45, 1459 (2005), ISSN 0042- 6989. Vision Res 47, 1518 (2007), ISSN 0042-6989. [47] Z. Li, Neural Comput 13, 1749 (2001). [48] Z. Li, Neural Comput 10, 903 (1998). [49] T. N. Mundhenk and L. Itti, Biol Cybern 93, 188 (2005). [50] W. Li, V. Piech, and C. D. Gilbert, Neuron 50, 951 (2006). [51] S. Grossberg and E. Mingolla, Percept Psychophys 38, 141 (1985). [24] O. Ben-Shahar and S. Zucker, Neural Comput 16, 445 [52] S. Ullman, R. L. Gregory, and J. Atkinson, Philos T R (2004). Soc Lon B 337, 371 (1992). [25] W. S. Geisler and J. S. Perry, Visual Neurosci 26, 109 [53] S.-C. Yen and L. H. Finkel, Vision Res 38, 719 (1998), (2009). ISSN 0042-6989. [26] S. Mandon and A. K. Kreiter, Vision Res 45, 291 (2005), [54] P. J. Garrigues and B. A. Olshausen, in Adv Neur In ISSN 0042-6989. (2007). [27] M. Ursino and G. E. L. Cara, Neural Networks 17, 719 [55] A. D. Ing, A. J. Wilson, and W. S. Geisler, J Vision 10, (2004). 1 (2010). [28] A. Sterkin, A. Sterkin, and U. Polat, J. Vis. 8, 1 (2008). [29] W. Bair, J. R. Cavanaugh, and J. A. Movshon, J Neurosci [56] H. Yao, L. Shi, F. Han, H. Gao, and Y. Dan, Nat. Neu- rosci. 10, 772 (2007). 23, 7690 (2003). [57] T. Hua, P. Bao, C.-B. Huang, Z. Wang, J. Xu, Y. Zhou, [30] N. R. Zhang and R. von der Heydt, J. Neurosci. 30, 6482 and Z.-L. Lu, Curr Biol 20, 887 (2010). (2010). [58] W. Li and C. D. Gilbert, J. Neurophysiol. 88, 28462856 [31] L. Schwabe, K. Obermayer, A. Angelucci, and P. C. (2002). Bressloff, J. Neurosci. 26, 9117 (2006). [59] A. Knoblauch and F. T. Sommer, Neurocomputing 58- [32] A. Angelucci, J. B. Levitt, E. J. S. Walton, J.-M. Hupe, J. Bullier, and J. S. Lund, J. Neurosci. 22, 8633 (2002). [33] T. Serre, A. Oliva, and T. Poggio, P Natl Acad Sci USA 60, 185 (2004). [60] P. O. Hoyer and A. Hyvarinen, Vision Res 42, 1593 (2002). 104, 6424 (2007). [61] S. Song, K. E. Miller, and L. F. Abbott, Nat Neurosci 3, [34] S. Martinez-Conde, S. L. Macknik, X. G. Troncoso, and D. H. Hubel, Trends Neurosci 32, 463 (2009), ISSN 0166- 2236. [35] E. T. Rolls and M. J. Tovee, P Roy Soc Lond B Bio 257, 9 (1994). [36] F. Wilkinson, H. R. Wilson, and C. Habak, Vision Res 919 (2000). [62] Z. Li, Trends Cogn Sci 6, 9 (2002), ISSN 1364-6613. [63] K. Fukushima, Biol Cybern 36, 193 (1980). [64] C. D. Gilbert and M. Sigman, Neuron 54, 667 (2007). [65] D. Schneeweis and J. Schnapf, Science 268, 1053 (1995). [66] J. T. Enns and V. D. Lollo, Trends Cogn Sci 4, 345 38, 3555 (1998), ISSN 0042-6989. (2000), ISSN 1364-6613. [37] W. S. Geisler, J. S. Perry, B. J. Super, and D. P. Gallogly, [67] L. Fei-Fei, R. Fergus, and P. Perona, in CVPR 2004, Vision Res 41, 711 (2001), ISSN 0042-6989. Workshop on Generative-Model Based Vision (2004). [38] E. Schneidman, M. J. Berry II, R. Segev, and W. Bialek, [68] Y. LeCun, L. Bottou, Y. Bengio, and P. Haffner, in P Nature 440, 1007 (2006). [39] J. Shlens, G. D. Field, J. L. Gauthier, M. I. Grivich, D. Petrusca, A. Sher, A. M. Litke, and E. J. Chichilnisky, IEEE (1998), vol. 86, p. 2278. [69] D. H. Brainard, Spatial Vision 10, 433 (1997).
1206.0637
1
1206
2012-06-04T14:50:03
A Quantum-mechanical description of ion motion within the confining potentials of voltage gated ion channels
[ "q-bio.NC", "physics.bio-ph" ]
Voltage gated channel proteins cooperate in the transmission of membrane potentials between nerve cells. With the recent progress in atomic-scaled biological chemistry it has now become established that these channel proteins provide highly correlated atomic environments that may maintain electronic coherences even at warm temperatures. Here we demonstrate solutions of the Schr\"{o}dinger equation that represent the interaction of a single potassium ion within the surrounding carbonyl dipoles in the Berneche-Roux model of the bacterial \textit{KcsA} model channel. We show that, depending on the surrounding carbonyl derived potentials, alkali ions can become highly delocalized in the filter region of proteins at warm temperatures. We provide estimations about the temporal evolution of the kinetic energy of ions depending on their interaction with other ions, their location within the oxygen cage of the proteins filter region and depending on different oscillation frequencies of the surrounding carbonyl groups. Our results provide the first evidence that quantum mechanical properties are needed to explain a fundamental biological property such as ion-selectivity in trans-membrane ion-currents and the effect on gating kinetics and shaping of classical conductances in electrically excitable cells.
q-bio.NC
q-bio
To appear in Journal of Integrative Neuroscience 11 (2), June 2012. A Quantum-mechanical description of ion motion within the confining potentials of voltage gated ion channels Johann Summhammera, Vahid Salarib, Gustav Bernroiderc,1 a Atom Institute, Vienna University of Technology, Stadionallee 2, A-1020 Vienna, Austria b Kerman Neuroscience Research Center (KNRC), Kerman, Iran c Department of Organismic Biology, Neurosignaling Unit, University of Salzburg Hellbrunnerstr 34,A-5020 Salzburg, Austria Email: [email protected] Abstract Voltage gated channel proteins cooperate in the transmission of membrane potentials between nerve cells. With the recent progress in atomic-scaled biological chemistry it has now become established that these channel proteins provide highly correlated atomic environments that may maintain electronic coherences even at warm temperatures. Here we demonstrate solutions of the Schrodinger equation that represent the interaction of a single potassium ion within the surrounding carbonyl dipoles in the Berneche-Roux model of the bacterial KcsA model channel. We show that, depending on the surrounding carbonyl derived potentials, alkali ions can become highly delocalized in the filter region of proteins at warm temperatures. We provide estimations about the temporal evolution of the kinetic energy of ions depending on their interaction with other ions, their location within the oxygen cage of the proteins filter region and depending on different oscillation frequencies of the surrounding carbonyl groups. Our results provide the first evidence that quantum mechanical properties are needed to explain a fundamental biological property such as ion-selectivity in trans-membrane ion-currents and the effect on gating kinetics and shaping of classical conductances in electrically excitable cells. Keywords: Quantum mechanics; quantum biology; quantum oscillations; Schrodinger equation; neural signaling; molecular dynamics; ion channels; coherence dynamics; selectivity filter; filter gating. 1 Introduction Ion channels are the building blocks of electrical membrane signals in the nervous system. They are responsible for the controlled charge transition across cell membrane, building up the trans- membrane potentials that propagate along the membranes as either dendritic potentials or action potentials. The ’channels’ are proteins inserted into the cell membrane and host different sub- domains that operate at highly different physical action orders, ranging from the quantum-scale at the single atom level up to the classical scale of the entire protein [4, 5]. The main parts of the protein consist of the pore domain gate, controlling the access of ions into the protein, the cavity region that hosts a hydrated ion prior to its access to the narrow ’selectivity filter’ that provides the last steps of charge transfer into the cell or outside the cell (see Figure 1). The selectivity filter (SF) of ion channel proteins is responsible for the selective and fast conduction of ions across neuronal cell membranes. After the determination of the atomic resolution structure of the bacterial model KcsA channel by MacKinnon and colleagues [15, 1], it became increasingly apparent that delicate atomic interactions within the highly ordered conduction pathway of the SF are critical for the ability of the protein to provide a very high rate of conduction without loss of selectivity for a particular ion species. Following these initial findings a long list of molecular dynamics studies (MD) added important functional aspects to the static details of the channel 1corresponding author 1 protein as reviewed by [2, 10, 14, 18]. Density functional calculations and hybrid QM/MM methods have substantially contributed in capturing the interaction between carbonyls and their effect on backbone atoms [8], together with a significant charge-polarization effect of ions on their filter ligands [12]. Taking into account these results, additional theoretical treatments of ion complexation in water and diverse binding site models have revealed the topology of forces that are engaged in providing a highly selective coordination of ions in the filter region [6]. The question became dominant how selectivity can emerge without compromising conduc- tion. Within this context at least two sets of problems are at the center of interest. First, it can be expected that the Coulomb type forces among ions , the adjacent water dipoles and the neighboring oxygen atoms, with average distances of 0,285 nm in the KcsA filter [13], require a quantum description, because the involved distances are near the de Broglie wavelength associ- ated with the ions at thermal energies. Are there typical quantum interaction effects or quantum interferences that are indispensible to explain the biological functioning of these channel proteins ? The second point involves the question, can quantum effects propagate into the classical states of proteins ? The last question gains considerable significance from findings about different ’gat- ing mechanisms’ of the channel. Whereas ’pore gating’ is a relatively slow process that controls the access of ions to the cavity region of the filter, the filter itself is found to be able to change between ’permissive’ and ’non-permissive’ conformational states at a much shorter time-scales [3, 18]. Because ’selectivity’ seems to be organized within the filter that is located opposite the ’entrance’ of the proteins pore region, it can be expected that the mechanisms providing selec- tivity must somehow become coordinated with the mechanisms controlling access of ions to the cavity region. If quantum effects do play a critical role for filter-ion coordination, it is feasible that these delicate interactions could leave their quantum traces in the overall conformation and the molecular gating state of the entire protein. In the present paper we investigate this question. We study the question if and how quantum interferences in the filter region of a typical K-channel can influence the atomic environment within the frame set by time-dependent potentials derived from carbonyl charges, water molecules and other ions present in the atomic environment of the filter-region. By solving the Schrodinger equation for the quantum mechanical states of K + ions, we demonstrate that, depending on the size of the confining potential and the thermal energy of the ion, the ion’s wave-function can become highly delocalized, with its probability distribution extending over a significant fraction of the filter region. This in turn exerts an influence on the motion and induced polarization of the carbonyl groups lining the filter in a way that is different from the interaction that could be expected by a strictly localized classical ion. Due to the interaction of the ions with a time- dependent potential, the energy of ions is not conserved. Instead, we observe tera-hertz (THz) oscillations of ionic wave-packets that become damped, giving off their energy to the environment either via the vibrational modes of the surrounding carbonyl dipoles or by radiation, or both. These effects will cool down the ions in the filter domain dramatically. We finally demonstrate the result of a ’replacement study’, where K + ions in the same environment are replaced by Na+ ions. The results show that the oscillation frequency of the coordinating carbonyl dipoles discriminates the kinetic energy minima for both ion species. This is the first indication of a quantum mechanical effect on the ion selectivity function of voltage gated channel proteins. 2 Methods 2.1 The molecule We represent the selectivity filter as an array of C=O dipoles provided by the backbone carbonyl groups of the highly conserved TVGYG amino acid sequence (see Figure 1,2). Each filter segment carries five dipoles. We account for the intermittent, nearest neighbor water molecules in the filter 2 Figure 1: The main physical operation facilitated by an ion channel-protein is an ion-species selective charge transfer (A), where one charge is taken up from the cavity (shown left), dehy- drated and transferred into the filter region (marked by double bars). The filter region hosts at least two ions. The uptake of one ion leads to a knock-on effect on the ions present in the filter region. One ion is taken up from the cavity site, and another ion is released to the (intra or extra) cellular site (shown right). (B) The basic organization of voltage gated ion channels. The protein consists of three subdomains, the pore, the cavity and the filter region. 3 Figure 2: A schematic illustration of a dimeric KcsA view with filter location and associated trans-membrane helices. The carbonyl atoms are indicated by short segments and numbered S0-S4 starting from the extracellular site. In (B)a demonstration of a free energy distribution given in kcal/mol along the axial z-coordinates (in Angstroems) of the filter according to the Berneche-Roux model is shown. The ion-filter configuration for the energy distribution in (B) is demonstrated by the scheme(C). In the present paper the ion at location S1 (insert C) is considered quantum-mechanically, the surrounding atoms from carbonyl groups, other ions in the filter region and the intermittent water molecule are treated as classical particles. by a confining potential for the ions. This water-derived potential behaves similar to repelling atoms, degrading the spread of the ion’s wavepacket. The distances of carbonyl C-atoms to the filter axis were preset to 0,345 nm according to the Garafoli-Jordan permeation model [11]. Effective charge distances of C=O dipoles were set to 0.122 nm. Bond directions along C=O groups are variable with zero degrees direction perpendicular to the filter axis. In the present hybrid model we consider one potassium ion as a quantum particle and treat the surrounding nuclear motion classically. 2.2 The quantum model The basic interaction between classical ions and C=O dipoles as well as a quantum-mechanically described ion and it’s atomic surrounding is the electrostatic repulsion and attraction. As usual the potential between two electrical point charges is V (r) = 1 4πǫ · q1q2 r (1) We have used ′r′ for the distance between two particles and ’q’ for their charges. The symbol ′ǫ′ denotes the dielectric constant of the medium containing the charges. Taking into account a total of 20 C=O dipoles the resulting potential for the region containing one K + ion becomes: VF (r, t) = qion 4πǫ 20 Xj=1   (cid:12)(cid:12)(cid:12) ROj (t) − r(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12) RCj (t) − r(cid:12)(cid:12)(cid:12) Here the vector ¯r represents the ions position at time t, qOj for the charge of the j-th oxygen qOj qCj +   (2) 4 and qCj for the charge of the j-th carbon atom respectively. Further ¯ROj denotes the time- dependent position of the j-th oxygen and ¯RCj the position of the j-th carbon atom. In additon we have to consider the potential Vi that is contributed by the classical ions in the channel: VI (r, t) = qion 4πǫ   (cid:12)(cid:12)(cid:12) RK1(t) − r(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12) RK2(t) − r(cid:12)(cid:12)(cid:12) qK qK +   (3) The potential due to the water molecules on both sides of the ion is given by a potential VH2O, which strongly depends on the form of the wave-packet. The time-dependent locations of classical ions have been denoted by RKj . The total potential to which the quantum particle is exposed is thus: V (¯r, t) = VF (¯r, t) + VI (¯r, t) + VH2O(¯r, t) (4) From a modelistic view the overall axial symmetry of the channel (Figure 3) allows for a restriction of the calculations to one spatial dimension, along the z-axis of the filter. This also reduces the computational effort considerably. Within the frame set we obtain a Schroedinger equation for one K + in the filter: −¯h2 2mk d2 dz2 ψ(z, t) + V (z, t).ψ(z, t) = i¯h ∂ ∂t ψ(z, t) (5) The water molecules interacting with the ion move with the ion and thus somewhat impede the spreading of its wave-packet by an effective potential: VH2O(z, t) = −EH2O.nZ d −d ψ(z + s, t)2 ds (6) Here, EH2O denotes the depth of the confining potential, set to 10 meV, which corresponds to about half the thermal energy kBT . The symbol ’d’ represents the average distance between the oxygen of a water molecule and the K + ion as available from both, experimental [16, 26] and theoretical studies [26]. For the present study we have set a d = 0.28 nm. Further ’n’ is a normalization constant which ensures that the minimum of VH2O(z, t) equals −EH2O. We have calculated the wave-function together with the resulting space and time dependent probability p = ψ(z, t)2 for different oscillations of C=O groups and locations of classical ions in the filter region. All second order differential equations were solved by applying the Crank-Nicolson formalism [23]which is known for it’s stability over a large range of calculational steps. Additional controls were applied to check the consistency of calculations by doubling the temporal resolution within a constant time interval (e.g. 678.500 steps with 0.006 fs to 2.750.000 steps with 0,0015 fs resolution, see Figure 4). No numerical artifacts on the stability of the results could be observed. Also, under constant, time independent potentials the wave-packets behavior was as can be expected for harmonic oscillating potential forms, for which analytic solutions are well known. The initial wave-function was approximated by a Gaussian distribution centered at one of the possible filter locations (S0-S4). 3 Results 3.1 Conductive States In the following figures we provide some snapshots of varying potentials confining the quantum- mechanical ion KQ at location S1 of the filter region for different interactions with ’classical ions’ KB and KA moving into the filter region. The filter sets out with a double occupancy of 5 Figure 3: A three-dimensional distance model for the carbonyl-ion coordination within the filter region of the KcsA model channel as used for the present study. The 20 C=O dipoles are symbolized by circles and are numbered on the left to specify the distance matrix that we used for the present calculations and the 3-dim topology provided by the right-hand figure (dark lines join the C=O dipoles from two lateral strands and white lines join opposing C=O dipoles) Figure 4: Calculational consistency of the Crank-Nicolson approximation: a comparison of wave- function solutions with different temporal scales, ranging from 687.500 steps with a time width of 0.006 fs (top) to 1.375.000 steps (at 0.003 fs) (middle) and 2.750.000 steps (at 0.0015 fs) (bottom). Changes in the probability distribution of the ions spatial location (red curve at the bottom of the inserts) are not detectable, demonstrating the robustness and stabilitiy of the Crank-Nicolson method for solving equations of the type ut = a.uxx (e.g. the Schrodinger equation) as used in the present study. 6 Figure 5: QM-wave-packet distribution of one ion centered at location S1 interacting with two classical ions KA and KB that enter the filter region. The orientation of the filter atoms is from inside the cell (left) to outside the cell (right) according to the scheme shown in Figure 2(C). The wave-packet distribution of the ion KQ at location S1 is shown within dark stripes in order to visualize the emerging interference (coherence) pattern of its wavefunction. The confining potentials within these stripes are demonstrated in the following Figure 6 for four different time step situations, labeled by (a-d). The circles next to the ions, with two rows above and two rows below lining the filter, symbolize the location of C atoms (outside) and O atoms (inside) that belong to the carbonyl C=O groups of the filter. Note the distortion of C=O dipoles induced by the presence of the KQ ion in S1 ions at location S1-S3 and develops into ion configuration S2-S4 with the outermost quantum- mechanically calculated ion close to S0. The snapshot shown in the following Figure 5 and Figure 6 cover a time-window of just 4.124 ps. The time sequence of frames (a-d) in Figure 5 and their matching potentials (a-d) in Figure 6 cover the ’conductive’ situation with one ion (KB) approaching the filter region from the cavity site, another ion (KA) located at S3 and the wave-packet of the ’quantum-ion’ (KQ) centered at S1. During this time (KQ) oscillates as a wave around it’s center due to the thermal energy of the ion. In Figure 2 (b) the external ion exerts a repulsive ’knock-on’ effect on the (K A). This external knock-on is also felt by the quantum-ion ( K Q) as it now extends it’s wave-function with some probability spreading into location S0. This non-local effect covers a z-extension in the order of 10−10m. Finally, as seen in the frame Figure 5(d) the classical filter ion has reached position S2 inducing a strong non-local effect on the spread of the wave-function of the quantum ion, originally located at S1. There is now a considerable probability extension of 1nm (about the entire linear z-extension of the filter) for the amplitude of ion KQ, with a spread into S0 and into the extra-cellular site of the filter. 7 Figure 6: Reconstructed dynamics of potentials along the filter locations S1 and S0 that results from the movement of 2 (classical) ions KA, KB and one ’quantum-mechanical ’ ion KQ during four consecutive snapshots (a-d) with locations as shown in Figure 2C. The potentials are shown within the range of 0.1 to 0.5 eV. The x-axis is provided by the z -coordinates of the filter region as shown in Figure 2B. It should be noted that the confining potentials for the ions are not pre-set, but the potential distribution emerges from the Coulomb interaction of ions with the charges provided by the lining oxygen atoms and the other ions present in the filter following the calculations by equations (2.2 - 2.4) 8 Q i ion and a N a+ Figure 7: Mean kinetic energy of a K + Q ion as a function of oscillation frequencies of filter C=O groups during 9 ps of simulation. The replacement of the ’intrinsic’ K + Q ion with a N a+ Q ion with different coordination distances to the surrounding atoms becomes reflected by a different kinetic energy minimum along the oscillation frequency of carbonyl dipoles. The loss of kinetic energy (’cooling’) is less for the replaced atom species and requires a different oscillation frequency of C=O groups. 3.2 Non-conductive states For the non-conductive states of the filter we assume that the classical ions occupy the cavity position and the filter position S3 initially. The quantum-ion remains in position S1 (the so-called S1-S3 filter configuration). We now allow the interacting C=O groups to oscillate thermally. This will in turn induce an oscillation of local potentials around the quantum-ion at S1. Cooling effects: It can be expected that the translocation of ions, followed by a transient binding into a local potential, will generally lead to a loss of energy of this ion and a subsequent ’cooling effect’. In an attempt at understanding the role of the filter structure for these possible effects, we have made a series of measurements on the dependence of mean kinetic energies on different oscillation frequencies of the coordinating C=O groups. The initial settings placed the (KQ) ion into S1 with a kinetic energy of 20 meV (1 kT = 26 meV at room temperature), the wave-packet extension was preset to 0.025 nm and the oscillatory amplitude of C=O to 0.02 nm. We followed the evolution of the K Q wave-function during a period of 9 ps at intervals of 0.006 fs (over 1.5 million steps). In Figure 7 we show the mean kinetic energy of this ion at the end of the 9 ps cooling period as a function of the oscillation frequency of the C=O groups. The temporal development of the kinetic energies over the time window of 9 ps is demonstrated in Figure 8. From Figure 8 it is apparent that at optimal oscillation frequencies of C=O groups (open circles in Fig.7) the kinetic energy approaches a minimum after a quarter cycle of the oscillation period. At this point the potential energy reaches a maximum. However, because the potential sink follows the movement of the ion, this potential energy is only partly returned into the kinetic 9 Figure 8: The temporal evolution of kinetic energies of the K + Q ion as shown in Figure 7 at site S1 for three different oscillation frequencies: unbroken line = no oscillations, broken line = 588 GHz oscillations and circles = the ’optimal frequency’ at 446 GHz with maximal cooling effects. The time evolution of these curves shows that only for the oscillation frequency that provides a clear cooling effect (Figure 7) at 446 GHz, the potential energy is not returned to the kinetic part during the next oscillation period. Thus the concerted quantum-mechanical behavior of the ion together with the coordinating carbonyl dipoles can account for this unique selection procedure at the atomic scale. 10 term during the next quarter of the oscillation. This process continues until the kinetic energy of the ion approaches a minimum, depending on the frequency of the associated C=O oscillations. Further, particular oscillation frequencies of the ion-carbonyl environment and the associated energy dissipation can discriminate different ion species for a given atomic filter environment. Whereas a primary ion selection effect seems to be associated by different ion-species dependent dehydration energies during the ions transfer from the cavity to the selectivity filter [17], the present findings are strongly suggestive for an additional role of quantum oscillatory effects within the filters atomic environment on ion selectivity. This points to a clear quantum effect that is indispensible in explaining ion selectivity, a basic and indispensible biological function of ion channel proteins. As expected, our results show that if an N a+ ion is placed in the same position as a K + , it also can transfer energy to the C=O lattice (see Figure 7). However, since the N a+ ion has a lower mass than the K + ion, the most efficient cooling occurs at a higher oscillation frequency of the C=O groups. The natural oscillation frequency of the C=O groups caused by thermal motion of the surrounding atoms is essentially determined by the binding strength of the C=O groups to the back- bone of the protein strands. We can therefore expect this oscillation to exert its cooling effect in the most efficient way for ions with a specific charge and a specific mass. When hopping from one site to a neighboring site in the selectivity filter only a particular ion species will be captured optimally by the coordinating C=O cage. However, because the exchange of energy works both ways, the con-specific ion will also be able to extract energy from the coordinating C=O group within a shorter time than any other ion species. This in turn will allow shorter times for bridging the energy barrier to the neighboring site. Further simulations will be required to test whether this effect can lead to a possible resonance phenomenon accompanying a fast passage through the selectivity filter. 4 Conclusions We have shown that, depending on the surrounding carbonyl derived potentials, alkali ions in the filter region of ion channels lined by the conserved TVGYG amino acid sequence can become highly ’quantum-delocalized’ depending on ion location and oscillatory motion of the coordinating C=O groups. Further, we find that due to the interaction of ions with a time-dependent potential, the energy of ions is not conserved. Depending on the frequencies of oscillating C=O surroundings the ion can lose up to 1/2 of its kinetic energy, exerting a substantial ’cooling effect’. The transfer of this energy to the C=O environment is shown to depend on a specific C=O resonant oscillation frequency and within the present study is observed under a ’non-conductive’ filter state for an ion at location S1. In the frame of previous work on energy transfer systems in proteins, quantum and classical [7, 20, 21]we suggest that the observed effect will cool down the ions inside the filter giving off their energy to the environment, i.e. to the so-called P-loop domain of the lining filter region. Energy transfer along the P-loop backbone atoms can in turn change the symmetry of the filter entrance [3, 9, 13]. As a result it was anticipated that even the location of a single ion in the filter can become associated with the gating state of the selectivity filter [3, 9]. Here we provide evidence for this situation based on a quantum-mechanical calculation and offer predictions about the involved energies and resonant oscillations and their effect on channel filter states. These results indicate that temporal development of quantum states in channel proteins can propagate into classical ion-channel conformations that determine the electrical signal properties of neuronal membranes. Because the filter states must be correlated with the pore-domain gating state of the channel in order to facilitate the co-occurrence of conduction and access of ions to the channel protein [24], the pore gating states of the channels can be interpreted as ’classical witness states’ [25] of an underlying quantum process in the brain, as suggested previously [4, 5, 19]. 11 References [1] Berneche S, Roux B, Energetics of ion conduction through the K channel, Nature 414: 73-77, 2001. [2] Berneche S, Roux B, A microscopic view of ion conduction through the K channel, PNAS 100: 8644-8648, 2003. [3] Berneche S, Roux B, A gate in the selectivity filter of potassium channels, Structure 13:591- 600, 2005. [4] Bernroider G, Roy S, Quantum-classical correspondence in the brain: scaling, action dis- tances and predictability behind neural signals, Forma 19: 55-68, 2004. [5] Bernroider G, Roy S, Quantum entanglement of K + ions, multiple channel states, and the role of noise in the brain, SPIE 5841: 205-213, 2005. [6] Bostick DL, Brooks III CL, Selectivity in K channels is due to topological control of the permeant ion’s coordinated state, PNAS 104:9260-9265, 2007. [7] Briegel HJ, Popescu S, Entanglement and intra-molecular cooling in biological systems? - A quantum thermodynamic perspective, arXiv :0806.4552v2 [quant-ph], 2009. [8] Bucher D, Raugei S, Guidoni L, Dal Peraro M, Rothlisberger U, Carloni P, Klein ML, Polarization effects and charge transfer in the KcsA potassium channel, Biophys.Chemistry 124:292-301, 2006. [9] Cordero-Morales JF, Cuello LG, Perozo E, Voltage-dependent gating at the KcsA selectivity filter, Nature Structural & Molecular Biology 13:319-322, 2006. [10] Corry B, Understanding ion channel selectivity and gating and their role in cellular sig- nalling, MolBioSyst2:527-535, 2006. [11] Garofoli S, Jordan PC, Modelling permeation energetics in the KcsA potassium channel, Biophys J 84:2814-2830, 2003. [12] Guidoni L, Carloni P, Potassium permeation through the KcsA channel: a density functional study Biochim Biophys Acta 1563:1-6, 2002. [13] Kona J, Minozzi M, Torre V, Carloni P, A gate mechanism indicated in the selectivity filter of the potassium channel KscA, Theor Chem Acc 117:1121-1129, 2007. [14] Kuyucak S, Andersen OS, Shin-Ho Chung, Models of permeation in ion channels, Rep Prog Phys 64: 1427-1472, 2001. [15] MacKinnon R, Potassium channels, FEBS Lett 555:62-65, 2003. [16] Mahler J, Persson I, A study of the Hydration of the Alkali Metal Ions in Aqueous Solutions, Inorg Chem (online pre publication),2011. [17] Rempe SB, Pratt LR, Ion Hydration Studies Aimed at Ion Channel Selectivity, Los Alamos, Special Feature April 2000, Theoretical Chemistry & Molecular Physics T-12, 2000. [18] Roux B, Schulten K, Computational studies of membrane channels, Structure 12:1343-1351, 2004. 12 [19] Roy S, Llinas R, Relevance of quantum mechanics on some aspects of ion channel function, C.R. Biologies 332:517-522, 2009. [20] Salari V, Summhammer J, Bernroider G, Coherent excitation energy transfer (EET) along the P-loop of KcsA ion channel models: Feasibility and possible effects on filter gating, QuEBS2010, Abstracts, 2010. [21] Salari V , Tuszynski J, Rahnama M, Bernroider G, Plausibility of quantum coherent states in biological systems, J Phys: Conf Series 306:012075, 2011. [22] Soper AK, Weckstrom K, Ion Solvation and Water Structure in Potassium Halide Aqueous Solitions. Biophys Chem 124: 180-191,2006. [23] Thomas JW, Numerical Partial Differential Equations: Finite Difference Methods, in Text- books in Applied Mathematics, 22, Berlin, New York , 1995. [24] VanDongen AM, K channel gating by an affinity-switching selectivity filter, PNAS 101: 3248-3252, 2004. [25] Vedral V, Quantifying entanglement in macroscopic systems, Nature 453:1004-1007,2008. [26] Zhou Y, Morais-Cabral JH, Kaufman A, MacKinnon R, Chemistry of ion coordination and hydration revealed by a K + channel-Fab complex at 2.0 A resolution, Nature 414: 43-48, 2001. 13
1711.01423
1
1711
2017-11-04T10:45:14
What can neuronal populations tell us about cognition?
[ "q-bio.NC" ]
Nowadays, it is possible to record the activity of hundreds of cells at the same time in behaving animals. However, these data are often treated and analyzed as if they consisted of many independently recorded neurons. How can neuronal populations be uniquely used to learn about cognition? We describe recent work that shows that populations of simultaneously recorded neurons are fundamental to understand the basis of decision-making, including processes such as ongoing deliberations and decision confidence, which generally fall outside the reach of single-cell analysis. Thus, neuronal population data allow addressing novel questions, but they also come with so far unsolved challenges.
q-bio.NC
q-bio
What can neuronal populations tell us about cognition? Iñigo Arandia-Romero1, Ramon Nogueira1, Gabriela Mochol1, and Rubén Moreno-Bote1,2 1 Center for Brain and Cognition & Department of Information and Communications Technologies, University Pompeu Fabra, 08018 Barcelona, Spain 2Serra Húnter Fellow Programme, 08018 Barcelona, Spain Corresponding author: Rubén Moreno-Bote Email: [email protected] Contributions: All authors discussed and wrote the paper Acknowledgements: This work has been supported by the Spanish PSI2013-44811-P and SlowDyn FLAGERA-PCIN-2015-162-C02-02 grants from MINECO to R. M. B., and IJCI- 2014-21937 grant from MINECO to G. M. Highlights  Neuronal population (NP) level analysis is slowly pervading systems neuroscience  NPs provide both unprecedented temporal and spatial resolution to study decision-making  However, large NPs come with challenging data analysis and interpretative problems  Challenges can be alleviated by using models of NP activity Abstract Nowadays, it is possible to record the activity of hundreds of cells at the same time in behaving animals. However, these data are often treated and analyzed as if they consisted of many independently recorded neurons. How can neuronal populations be uniquely used to learn about cognition? We describe recent work that shows that populations of simultaneously recorded neurons are fundamental to understand the basis of decision-making, including processes such as ongoing deliberations and decision confidence, which generally fall outside the reach of single-cell analysis. Thus, neuronal population data allow addressing novel questions, but they also come with so far unsolved challenges. Introduction Single-neuron electrophysiology has provided golden ages in neuroscience passing from the discoveries of Lord Adrian [1] to those by Hubel and Wiesel [2]. This approach relies on finding mappings between the activity of a single neuron and external world, body or internal variables, most often summarized with the so-called 'tuning curve', that is, the mean neuron response as a function of the relevant variable. As such, single-cell electrophysiology has proven to be a central pillar over which many theories of the brain rest [2-8]. In the last 20 years, there has been a surge of new technologies that allow recording the activity of hundreds of cells within and across brain areas, either with dense electrodes or with imaging techniques [9-16] and pioneering work has shown the benefits of these neuronal population (NP) recordings [17-22]. New emerging technologies even promise several orders of magnitude improvement in the number of recorded cells, leaping from hundreds to thousands of cells [23, 24]. However, these promises come with new problems. The most prominent challenges arise in experimental settings that involve animal behavior, where the number of trials is limited to avoid animal satiation and exhaustion, and to minimize overtraining to better study naturalistic behaviors. Considering the activity of each neuron as one dimension in the neuronal state space, how should we analyze and model multi- dimensional neuronal activity with just a few tens of trials per condition? How can we benefit from the many dimensions that NP offers? And, most importantly, what are the new questions that can be addressed with NPs about cognitive processes that cannot, or at least hardly, be asked with single neurons? We review recent literature that can help to answer the above questions with the focus on decision-making. We provide examples of a better understanding of the transformation from stimulus to choices by using the activity of simultaneously recorded cells. We argue about the difficulty, and even impossibility, of drawing qualitatively similar conclusions using single-cell electrophysiology. Then, we talk about the challenges that we face with large NPs and limited number of trials, as those expected from experiments that combine both animal behavior and NPs recordings. Finally, we point to past and more recent work that could benefit from using NP recordings and discuss possible promising future directions. Advantages of NPs The success of NP approaches has been clearly exemplified in brain-machine interface applications [25-28]. Recent studies have shown how it is possible to restore hand control in a quadriplegic patient [29] or aid non-human primates in walking [30] by reading out the population activity of motor cortex and converting these signals into motor commands bypassing injured nerves. 'Reading out' or 'decoding' means computing an online estimate of the value of an external world or internal variable from the activity of the recorded neurons [31-33]. For instance, decoding the population activity of V1 neurons with linear classifiers can predict with high accuracy the orientation of a stimulus drifting grating in a trial-by-trial basis [20, 34, 35]. Recent literature tends to favor approaches based on decoding over information theory [19, 20, 34, 36-43], mostly because decoding techniques can be applied with relatively small number of trials while information-theoretic methods require computing probability distributions over as many dimensions as neurons in the NP, a problem daunted by the curse of dimensionality [33]. In addition, decoding techniques provide trial-by-trial estimates of external and internal variables that can be directly interpreted and compared to their actual values [44]. Despite the success of NP approaches to understand the basics of attention [36], working memory [45] and decision-making [37, 38, 46], they are sometimes contemplated with suspicion. Indeed, in some applications it could be argued that NPs do not offer radically new knowledge compared to single-neuron approaches. For instance, while decoding analysis can provide better estimates of information content in a NP by taking into account the trial- by-trial correlated noise among cells, rough estimates can be obtained by extrapolating the information of singles neurons to small NPs of a few tens of cells [47] (but not for populations larger than 100-1000 cells due to the presence of 'differential correlations' [48]). Moreover, research based on NP recordings sometimes ends up using traditional single- neuron-based analysis [49, 50]. What are the unique opportunities that NPs can offer? Here we give examples of how NPs can provide new knowledge that is very hard - and in some cases impossible - to acquire from single-neuron-based analysis. There are at least two broad families of problems that cannot be addressed with single-neuron approaches: 1) internal neuronal dynamics that is not time-locked to observable stimulus or body variables, and 2) variables that are encoded at the population level and not at the single-neuron level. The first category relates to neuronal processing that is internally or externally triggered without experimental control. These phenomena might include trial- by-trial fluctuations of attention that are not cued experimentally, mind wandering, and any other neuronal processing that is not time-locked to the stimulus or body variables [34, 36, 37, 39, 41, 45, 51-63]. Because of their lack of significant correlation to observable experimental variables, these phenomena cannot be fully studied using single-neuron approaches. The second category relates to neuronal processing that requires coordination of neurons in a NP. For instance, stable representation of a world variable, such as world- centered body location of mice [64], can occur despite unstable representation of that variable at the single-cell level (Fig. 1a). Similarly, theoretical work shows that stable representation of working memory signals can be achieved in NPs despite unstable single- neuron representations [65]. Thus, observing single neurons can lead to the erroneous conclusion that the encoding of the variable is unreliable, while the reality is different, namely, the variable is reliably encoded at the NP level. Further, theoretical studies suggest that the encoding of variables can involve temporal delays between neurons without concomitant changes of activity in any single cell [66] (Fig. 1b), and conclusions on information content drawn from combining single cells recordings can be misleading if there is shared response variability among cells in the NP [32, 48] (Fig. 1c). Opportunities from NPs to understand decision-making NPs can provide new ways to study the processes underlying decision-making and cognition. These novel opportunities with NP arise because crucial decision-making variables might not be generally representable by single neurons. . An example of such a variable is 'decision confidence' [67-71], that is, the confidence of a decision maker about her choice being correct based on accumulated noisy evidence [68]. A large family of mathematical models for two-alternative decision-making tasks called 'race models' requires at least two types of neurons racing to generate the choice [67, 68]. Here each racing neuron represents at each time point the evidence for each option accumulated so far (Fig. 2). For this family of models, decision confidence is encoded in the difference between the activity of two racing neurons (in addition to elapsed time), the so-called 'balance of evidence', a quantity that cannot be represented by the activity of neither of the neurons alone [68, 72]. The activity of any single racing neuron has little information about confidence. For example, it is possible to have no change in firing rate with confidence (Fig. 2a,b, cell 1) or even increase in firing rate (Fig.2, cell 2) for lower confidence (lower panels of Fig 2a vs Fig. 2b). More realistic models based on probabilistic population codes [73], which incorporate the Poisson-like nature of neuronal responses, also represent both choices and confidence simultaneously in the difference in activity of different neuronal pools [74]. Recent experimental work has shed new light into how changes of mind might be encoded in neuronal populations during decision-making, a process that can also guide further studies in the encoding of decision confidence on NPs. Kiani and colleagues recorded the activity of large NPs in area 8Ar of prearcuate gyrus while macaque monkeys performed a motion discrimination task [37]. The authors used decoding techniques to find a hyperplane in multidimensional activity space that separated well animals' choices on a trial-by-trial basis (Fig. 3a). Importantly, the distance of the population activity vector from the hyperplane could be used to infer in which trials the animal 'changed its mind' (Fig 3b). This method allows studying the ongoing process of choice at unprecedented levels of temporal resolution. Decision confidence theories predict that confidence should be related to the same NP activity distance, but these predictions remain to be tested. A recent paper by Rich and Wallis has studied the deliberation process in economic- based choices at a surprising level of temporal resolution using up to 16 electrodes in areas 11 and 13 in the monkey orbitofrontal cortex [38]. Following a smart experimental design, a linear decoder learned to discriminate between different offers presented on forced-choice single-offer trials, showing how offers were encoded in the NP when they were presented separately. This decoder was then used without retraining to predict the internal deliberation process in free-choice trials where two offers were simultaneously available, uncovering alternations over time between the states representing the available options (Fig. 3c). The untested assumption in this work is that the representation of offers is the same in forced and free choice trials. However, consistently with this assumption, longer total duration of the state corresponding to the chosen offer was observed compared to the unchosen one. Thus, the activity of NPs reveals a rich dynamic representation of choice options, which would be very hard to obtain with single neurons. Similarly complex dynamics has been also reported in other tasks and areas during decision-making [41, 45]. Challenges from NPs The above examples illustrate the importance of NP analysis. But what is the NP size that we ought to record from to address a given question? A priori, there is no limit to the desired size: the larger, the better, although in some cases with just a bunch of simultaneously recorded cells it is possible to address questions that are unfeasible with just a single cell, as described above for relevant theories of decision confidence. As a rule of thumb, the larger the NP, the more precisely the state of the neuronal network can be assessed [34, 53, 75], the more accurately stimulus or body variables can be decoded [20, 34, 35], and the more variables can be studied simultaneously [29, 30]. However, the promises of population-based analysis do not come free of challenges. The most important challenge is model specification given large number of parameters and limited number of trials. Consider, for instance, a population of 100 neurons with 200 trials per distinct condition and a binary classification task (400 trials in total). Using a quadratic discriminant for this task requires estimating a number of parameters that scales as the square of the number of neurons, around 5 thousand parameters, which is outside the reach of this limited amount of data. A linear classifier could be still used in this condition because 400 trials is in general sufficient to estimate its 101 free parameters. However, consider now a population of 1000 neurons with again 200 trials per condition and the same binary classification task. In this case, even for a linear classifier, we need to estimate 1001 parameters, which is above the number of data points collected. Roughly speaking and based on our experience, we would need at least around 5 times more trials per class and per neuron recorded to train a linear classifier (5neuronsclasses), which for large NPs will certainly be difficult to achieve. The problem is even bigger for more complex decoders with many more free parameters, such as deep feedforward networks [76]. Deep learning approaches that have boosted machine learning in the last few years can be hardly suitable for neuronal data based on a few trials per neuron and condition [76]. Novel approaches that would work with standard 'few-trials' data need to be developed. What approaches are then suitable for 'few-trials' data? This is a domain ripe for research, and here we discuss two recently used methods. The first one is to use linear decoders together with regularization techniques [37, 52, 77]. In this way, parameters that provide none or little explanatory power are pushed to zero or very small values, therefore reducing the complexity of the decoder and thus dampening overfitting. Regularization has been shown to be useful for NPs approximately as large as 150 neurons with more than 500 trials per condition [37]. However, a possible danger is that regularization implicitly introduces priors, which constraint the family of solutions, but might turn to be wrong. The second method consists of building models with increasing complexity starting from basic principles and performing model comparison using cross-validation techniques [34, 78, 79]. Basic principles refer here to the knowledge that has been acquired across many years of research. For instance, knowledge about translational invariance of real-world objects incorporated into convolutional networks allows to significantly reduce their number of free parameters, thus improving dramatically their classification performance [76]. Another example of prior knowledge that can be included in decoders is the Poisson-like nature of neuronal responses [4, 73, 80-83] and the gain modulations induced by global fluctuations in population activity [34, 51, 54, 75, 78, 84]. For instance, a recent model of NP activity with Poisson-like firing and both heterogeneous multiplicative and additive gains has been shown to explain better macaque V1 responses than alternative models ('multi-gain' model [34]) (Fig. 4). The first important advance is that this analysis directly exploits NP data, as otherwise it is impossible to study the modulation of single-neuron activity with global gain. Secondly, the resulting model has culminated research that has shown that global gain modulations can be either purely additive [51, 75], purely multiplicative [84], homogeneous multiplicative and additive [54], or a mixture of homogeneous multiplicative and heterogeneous additive gains [78] (Fig. 4). Finally, it is important to avoid some potential pitfalls when using NPs. First, as larger NPs are available, computing 'noise correlation' matrices (correlations between all pairs of neurons at fixed stimulus conditions, [32, 85, 86]) will become unfeasible, because the covariance matrix becomes singular when the number of neurons exceeds the number of trials [87], unless strong priors are used [88]. There is also a tendency to build large surrogate NPs from independently recorded single-neurons [63, 89, 90]. Surrogate NPs can be useful in some special circumstances [63, 89-91], but overusing them could lead to erroneous conclusions. This can be the case when neuronal populations contain 'differential correlations', which can be very small and yet have a huge impact on information and computations, and are absent in surrogate populations [48]. Conclusions: what can NPs offer in the future? We have shown that NP analysis is slowly pervading neuroscience, but fulminating strides are expected to occur soon. First, ongoing debates on the role of certain brain areas in decision-making and on its neuronal mechanisms will dramatically benefit from measuring the activity of very large NPs [92-95]. Second, decision-making is characterized by complex behavior that in real-life conditions involve a myriad of variables, such as complex stimuli and body variables. The many-fold dimensions of neuronal activity will allow studying the simultaneous encoding of those variables in the brain. Dimensionality reduction techniques [96], including decoders, will leave room to new methods (e.g., [97]) more suitable when the complexity of the experimental design does not permit for such reductionist, low dimensional, view of neuronal activity, as computational neural network models predict [98- 100]. Finally, it is also expected that new-generation decoders and other approaches suitable for 'few-trials' datasets, typical for behavioral and systems neuroscience, will allow precise estimates of sensory information from increasingly large NPs [48, 86]. Figures and legends Figure 1: Neuronal populations (NPs) are necessary to understand the neuronal code. (a) NPs can represent stable information while the tuning of individual neurons changes [64]. The activity of cell 1 (red) and cell 2 (blue) for different positions can change from the first day (top-left) to the tenth day (top-right), while the information represented by the NP remains constant (bottom). The ordering of the cells by preferred position differs from the first (bottom-left) to the tenth (bottom-right) day due to changes in individual tuning, but jointly they represent the same map. To recover the information represented by the NP a readout neuron (black circle at the top) adapts the weights (𝑤𝑖-s) to the tuning of each neuron (red and blue circles) at a given day. The information recovered across days is the same. (b) Stimulus information in theory can be encoded in the relative spike timing between neurons [66]. Vertical bars represent the spiking activity of cell 1 (red) and cell 2 (blue) in three cases while two stimuli (orange and apple) are presented consecutively (top). The first two cases show the activity recorded from single cells. Here, the readout neuron (black circle on the right) is not able to recognize the stimulus that has been presented. In the third case, in contrast, the readout neuron can measure the difference in the timing of spikes from cell 1 and cell 2, and recover the stimulus by using the relative spike latencies. (c) Pair-wise correlations can help decoding [32]. Each plot shows the mean activity (crosses) and the variability in the response (standard deviation represented by clouds) of two neurons when two different stimuli are presented (orange and green). The responses in the left panel have circular clouds, meaning that the variability is the same in all directions, and therefore the correlation between the two neurons is zero for both stimuli. The panel in the right shows two correlated neurons with the same amount of variability as in the left panel, but spread along the same direction (ellipses). The red line marks the decoding boundary, dividing the space between the area which corresponds to the orange stimulus and to the green one. The clouds overlapping in the left panel mean that in some trials where the green stimulus has been presented the decoder will predict the orange one, and vice-versa. However, for the correlated activity case (right) most of the trials are correctly classified, as the clouds do not overlap. Therefore, predicting the performance of a neuronal code depends on neuronal correlations, a problem that lies beyond single-neuron electrophysiology. Figure 2: Decision confidence in a 'race model'. (a) A decision maker choses between two offers (green apple and orange), whose values are represented by the activity of two neurons (red and green lines), growing over time as more evidence is accumulated. The decision rule is based on what option depicts the highest activity at decision time (dashed vertical line). In this case, the choice was easy because the values of the offers were very distinct, and therefore the confidence on the decision (represented by the distance between the activity of the neurons) was high [68]. (b) When the two offers have similar values (red vs green apple), the choice is more difficult, and thus decision confidence (distance between the races) is lower. The estimation of confidence would not be possible if the two neurons were not recorded simultaneously. Comparing the high and low confidence cases (a-b) the activity of the neuron preferring the green apple (cell 1) is identical, and thus this neuron alone does not correlate with decision confidence at the time of the decision. Although the neuron that prefers the lower value offer changes its activity in the two cases, the direction of this change is opposite to the change in confidence (cell 2 has higher rate for the low confidence trial) and cannot give a good approximation of the latter. nor provide information about the state of the other racing neuron, making confidence estimation impossible from single neuron recordings. Figure 3: Decision making in neuronal response space. (a) The dynamics of decision making of an animal performing a two-alternative task can be visualized in the neuronal response space, that is, the space where each direction corresponds to the activity of a recorded neuron. At a given time window the population activity is a single point in this space (blue and red dots) A linear decoder is trained to find a hyperplane (in this case, a plane in three dimensions) which best separates the points belonging to the two classes (red for T1 choice, and blue for T2 choice). At the beginning of each trial (faded colors) the population activity provides weak evidence about stimulus conditions, and thus the points are close to the hyperplane. As more evidence is accumulated, the population activity diverges away from the hyperplane, making easer to predict the animal's choice. The distance between population activity and the classification hyperplane defines the decision variable (DV, black line). (b) The temporal dynamics of DV reveals 'changes of mind'. In a given trial, the DV crossed the ambiguity point equal to zero (indicated by an arrow), suggesting the presence of a change of mind, that is, an instance in which the hypothetical intended choice flips (two trials are shown). Tracking these changes of mind would not be possible with single-neuron analysis. Reproduced from [37] with the permission of the publisher. (c) Ongoing deliberation inferred NP activity in monkey orbitofrontal cortex. A linear decoder, trained in forced-choice trials, was used to predict choices between simultaneously presented offers in free-choice trials. The posterior probability derived from the decoder for chosen (red), unchosen (blue) and unavailable options (gray, average of both unavailable options), are shown in six different trials. This analysis allows tracking the ongoing deliberation processes at unprecedented temporal resolution. Reproduced from [38] with the permission of the publisher Figure 4: Population activity models based on Poisson-like firing and global gain modulation. A large variety of models have been recently used to characterize the activity of NP including the effect of the population activity on single neuronal activity in primary visual cortex (V1) [34, 75, 78, 84]. (a) Population activity fluctuates across trials for the same stimulus (left panels). The tuning of a V1 neuron (right) is modulated with population activity (called here global gain), which is estimated as the sum of spikes across all neurons excluding the activity from the neuron for which tuning is being characterized; thus, this analysis cannot be done with single neurons, it requires NPs. Trials were ranked from high (top left) to low (bottom left) global gain for each stimulus condition. The activity of the selected neuron (green spike trains) was averaged across either the top (red box) or bottom (pink box) 50th percentile of trials, and the averages were plotted as a function of stimulus orientation (rightmost panel). The tuning was modulated with global gain (red vs. pink lines). Points and error bars are mean responses and s.e.m., respectively; lines are von Mises fits. Reproduced from [32] with the permission of the publisher. (b) Global gains can affect the tuning of V1 cells in a multiplicative (top-left) and/or an additive (bottom-left) way. The modulatory effect of the population activity on the tuning curves is heterogeneous, with different multiplicative (top-right) and additive factors (bottom-right) for different neurons. ***: p < 0.001. Reproduced from [32] with the permission of the publisher.(c) Models of NP activity including global gains. Neurons are assumed to fire as independent Poisson processes (or variations of this, see [78, 84]) conditioned to gain factors and/or , which reflect modulatory factors in the NP. The mean firing rate of each cell i in the NP is given by the function 𝑖 . The function 𝑖 determines the tuning of the cells when the gain factors are zero. The gains can produce multiplicative effects on the tuning curves if 𝑖 is non-zero, and additive effects if 𝑖 is non- zero (see panel b left)). A recent model for NP activity suggests heterogeneous multiplicative and additive modulations (multi-gain model [34]). The multiplicative gain model assumes that there is not additive gain and that the multiplicative gain is identical for all neurons ( 𝑖 𝑖 for all i). The affine model allows for heterogeneous additive gains, but the multiplicative gain is still uniform for all neurons ( 𝑖 for all i). The multi-gain model permits heterogeneous multiplicative and additive gains (it also uses ). The multi- gain model provides better fits of V1 population activity than other models. Cross-validated log-likelihood is shown. References 1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. 13. 14. 15. 16. 17. Adrian, E.D. and Y. Zotterman, The impulses produced by sensory nerve endings: Part 3. Impulses set up by Touch and Pressure. J Physiol, 1926. 61(4): p. 465-83. Hubel, D.H. and T.N. Wiesel, Receptive fields, binocular interaction and functional architecture in the cat visual cortex. Journal of Physiology London, 1962. 160: p. 106-154. Wilson, F.A., S.P. Scalaidhe, and P.S. Goldman-Rakic, Dissociation of object and spatial processing domains in primate prefrontal cortex. Science, 1993. 260(5116): p. 1955-8. Shadlen, M.N. and W.T. Newsome, The variable discharge of cortical neurons: implications for connectivity, computation, and information coding. J Neurosci, 1998. 18(10): p. 3870-96. Desimone, R., Face-selective cells in the temporal cortex of monkeys. J Cogn Neurosci, 1991. 3(1): p. 1-8. Rizzolatti, G. and L. Craighero, The mirror-neuron system. Annu Rev Neurosci, 2004. 27: p. 169-92. O'Keefe, J. and J. Dostrovsky, The hippocampus as a spatial map. Preliminary evidence from unit activity in the freely-moving rat. Brain Res, 1971. 34(1): p. 171-5. Miyashita, Y., Inferior temporal cortex: where visual perception meets memory. Annu Rev Neurosci, 1993. 16: p. 245-63. Maynard, E.M., C.T. Nordhausen, and R.A. Normann, The Utah intracortical Electrode Array: a recording structure for potential brain-computer interfaces. Electroencephalogr Clin Neurophysiol, 1997. 102(3): p. 228-39. Kelly, R.C., et al., Comparison of recordings from microelectrode arrays and single electrodes in the visual cortex. J Neurosci, 2007. 27(2): p. 261-4. Blanche, T.J., et al., Polytrodes: high-density silicon electrode arrays for large-scale multiunit recording. J Neurophysiol, 2005. 93(5): p. 2987-3000. Csicsvari, J., et al., Massively parallel recording of unit and local field potentials with silicon-based electrodes. J Neurophysiol, 2003. 90(2): p. 1314-23. Stosiek, C., et al., In vivo two-photon calcium imaging of neuronal networks. Proc Natl Acad Sci U S A, 2003. 100(12): p. 7319-24. Shoham, D., et al., Imaging cortical dynamics at high spatial and temporal resolution with novel blue voltage-sensitive dyes. Neuron, 1999. 24(4): p. 791-802. Buzsaki, G., Large-scale recording of neuronal ensembles. Nat Neurosci, 2004. 7(5): p. 446-51. Kerr, J.N. and W. Denk, Imaging in vivo: watching the brain in action. Nat Rev Neurosci, 2008. 9(3): p. 195-205. Schoenbaum, G. and H. Eichenbaum, Information coding in the rodent prefrontal cortex. II. Ensemble activity in orbitofrontal cortex. J Neurophysiol, 1995. 74(2): p. 751-62. 18. Warland, D.K., P. Reinagel, and M. Meister, Decoding visual information from a 19. 20. 21. population of retinal ganglion cells. J Neurophysiol, 1997. 78(5): p. 2336-50. Nikolic, D., et al., Distributed fading memory for stimulus properties in the primary visual cortex. PLoS Biol, 2009. 7(12): p. e1000260. Graf, A.B., et al., Decoding the activity of neuronal populations in macaque primary visual cortex. Nat Neurosci, 2011. 14(2): p. 239-45. Aertsen, A.M., et al., Dynamics of neuronal firing correlation: modulation of "effective connectivity". J Neurophysiol, 1989. 61(5): p. 900-17. 22. 23. 24. 25. 26. 27. 28. 29. 30. 31. 32. 33. Riehle, A., et al., Spike synchronization and rate modulation differentially involved in motor cortical function. Science, 1997. 278(5345): p. 1950-3. Ahrens, M.B., et al., Brain-wide neuronal dynamics during motor adaptation in zebrafish. Nature, 2012. 485(7399): p. 471-7. Kim, T.H., et al., Long-Term Optical Access to an Estimated One Million Neurons in the Live Mouse Cortex. Cell Rep, 2016. 17(12): p. 3385-3394. Lebedev, M.A. and M.A. Nicolelis, Brain-machine interfaces: past, present and future. Trends Neurosci, 2006. 29(9): p. 536-46. Andersen, R.A., et al., Cognitive neural prosthetics. Trends Cogn Sci, 2004. 8(11): p. 486-93. Collinger, J.L., et al., High-performance neuroprosthetic control by an individual with tetraplegia. Lancet, 2013. 381(9866): p. 557-64. Sadtler, P.T., et al., Neural constraints on learning. Nature, 2014. 512(7515): p. 423- 6. Bouton, C.E., et al., Restoring cortical control of functional movement in a human with quadriplegia. Nature, 2016. 533(7602): p. 247-50. Capogrosso, M., et al., A brain-spine interface alleviating gait deficits after spinal cord injury in primates. Nature, 2016. 539(7628): p. 284-288. Abbott, L.F., Decoding neuronal firing and modelling neural networks. Q Rev Biophys, 1994. 27(3): p. 291-331. Averbeck, B.B., P.E. Latham, and A. Pouget, Neural correlations, population coding and computation. Nat Rev Neurosci, 2006. 7(5): p. 358-66. Quian Quiroga, R. and S. Panzeri, Extracting information from neuronal populations: information theory and decoding approaches. Nat Rev Neurosci, 2009. 10(3): p. 173- 85. **34. Arandia-Romero, I., et al., Multiplicative and Additive Modulation of Neuronal Tuning with Population Activity Affects Encoded Information. Neuron, 2016. 89(6): p. 1305-16. This work demonstrates the influence of the population activity on the tuning of individual monkey V1 neurons. Population activity enhances or reduces information in small neuronal ensembles, depending on whether neurons are multiplicatively or additively modulated by population activity. Berens, P., et al., A fast and simple population code for orientation in primate V1. J Neurosci, 2012. 32(31): p. 10618-26. 35. *36. Cohen, M.R. and J.H. Maunsell, A neuronal population measure of attention predicts behavioral performance on individual trials. J Neurosci, 2010. 30(45): p. 15241-53. **37. Kiani, R., et al., Dynamics of neural population responses in prefrontal cortex 38. indicate changes of mind on single trials. Curr Biol, 2014. 24(13): p. 1542-7. Rich, E.L. and J.D. Wallis, Decoding subjective decisions from orbitofrontal cortex. Nat Neurosci, 2016. 19(7): p. 973-80. *39. Tremblay, S., et al., Attentional filtering of visual information by neuronal ensembles 40. in the primate lateral prefrontal cortex. Neuron, 2014. 85(1): p. 202-15. Adibi, M., et al., Population decoding in rat barrel cortex: optimizing the linear readout of correlated population responses. PLoS Comput Biol, 2014. 10(1): p. e1003415. **41. Morcos, A.S. and C.D. Harvey, History-dependent variability in population dynamics 42. during evidence accumulation in cortex. Nat Neurosci, 2016. 19(12): p. 1672-1681. Sussillo, D., et al., Making brain-machine interfaces robust to future neural variability. Nat Commun, 2016. 7: p. 13749. 43. Montijn, J.S., et al., Population-Level Neural Codes Are Robust to Single-Neuron Variability from a Multidimensional Coding Perspective. Cell Rep, 2016. 16(9): p. 2486-98. Churchland, A.K. and R. Kiani, Three challenges for connecting model to mechanism in decision-making. Curr Opin Behav Sci, 2016. 11: p. 74-80. 44. *45. Stokes, M.G., et al., Dynamic coding for cognitive control in prefrontal cortex. 46. 47. Neuron, 2013. 78(2): p. 364-75. Nogueira, R., et al., Lateral orbitofrontal cortex anticipates choices and integrates prior with current information. Nat Commun, 2017. 8: p. 14823. Ecker, A.S., et al., Decorrelated neuronal firing in cortical microcircuits. Science, 2010. 327(5965): p. 584-7. **48. Moreno-Bote, R., et al., Information-limiting correlations. Nat Neurosci, 2014. 17(10): p. 1410-7. In this study, it is characterized the type of neuronal pairwise correlations that limit information. A particular pattern of trial-by-trial shared variability is generated when the input layer is noisy itself, the so-called differential correlations, which make information saturate for large neuronal ensembles. Hosokawa, T., et al., Single-neuron mechanisms underlying cost-benefit analysis in frontal cortex. J Neurosci, 2013. 33(44): p. 17385-97. Sul, J.H., et al., Distinct roles of rodent orbitofrontal and medial prefrontal cortex in decision making. Neuron, 2010. 66(3): p. 449-60. Arieli, A., et al., Dynamics of ongoing activity: explanation of the large variability in evoked cortical responses. Science, 1996. 273(5283): p. 1868-71. Kiani, R., et al., Natural grouping of neural responses reveals spatially segregated clusters in prearcuate cortex. Neuron, 2015. 85(6): p. 1359-73. Okun, M., et al., Diverse coupling of neurons to populations in sensory cortex. Nature, 2015. 521(7553): p. 511-5. 49. 50. 51. 52. 53. *54. Mochol, G., et al., Stochastic transitions into silence cause noise correlations in 55. 56. 57. cortical circuits. Proc Natl Acad Sci U S A, 2015. 112(11): p. 3529-34. Scholvinck, M.L., et al., Cortical state determines global variability and correlations in visual cortex. J Neurosci, 2015. 35(1): p. 170-8. Jezzini, A., et al., Processing of hedonic and chemosensory features of taste in medial prefrontal and insular networks. J Neurosci, 2013. 33(48): p. 18966-78. Ponce-Alvarez, A., et al., Dynamics of cortical neuronal ensembles transit from decision making to storage for later report. J Neurosci, 2012. 32(35): p. 11956-69. **58. Engel, T.A., et al., Selective modulation of cortical state during spatial attention. Science, 2016. 354(6316): p. 1140-1144. This work demonstrates the influence of the population activity on the tuning of individual monkey V1 neurons. Population activity enhances or reduces information in small neuronal ensembles, depending on whether neurons are multiplicatively or additively modulated by population activity. Abeles, M., et al., Cortical activity flips among quasi-stationary states. Proc Natl Acad Sci U S A, 1995. 92(19): p. 8616-20. Seidemann, E., et al., Simultaneously recorded single units in the frontal cortex go through sequences of discrete and stable states in monkeys performing a delayed localization task. J Neurosci, 1996. 16(2): p. 752-68. Kaufman, M.T., et al., Cortical activity in the null space: permitting preparation without movement. Nat Neurosci, 2014. 17(3): p. 440-8. 59. 60. 61. *62. Kaufman, M.T., et al., Vacillation, indecision and hesitation in moment-by-moment decoding of monkey motor cortex. Elife, 2015. 4: p. e04677. *63. Mante, V., et al., Context-dependent computation by recurrent dynamics in prefrontal cortex. Nature, 2013. 503(7474): p. 78-84. *64. Ziv, Y., et al., Long-term dynamics of CA1 hippocampal place codes. Nat Neurosci, 65. 66. 67. 2013. 16(3): p. 264-6. Druckmann, S. and D.B. Chklovskii, Neuronal circuits underlying persistent representations despite time varying activity. Curr Biol, 2012. 22(22): p. 2095-103. VanRullen, R., R. Guyonneau, and S.J. Thorpe, Spike times make sense. Trends Neurosci, 2005. 28(1): p. 1-4. Vickers, D., Decision processes in visual perception. 1979, New York: London Academic Press. 406. *68. Moreno-Bote, R., Decision Confidence and Uncertainty in Diffusion Models with Partially Correlated Neuronal Integrators. Neural Computation, 2010. 22(7): p. 1786-1811. *69. Kiani, R. and M.N. Shadlen, Representation of confidence associated with a decision by neurons in the parietal cortex. Science, 2009. 324(5928): p. 759-64. *70. Kepecs, A., et al., Neural correlates, computation and behavioural impact of decision 71. 72. confidence. Nature, 2008. 455(7210): p. 227-31. Drugowitsch, J., et al., The cost of accumulating evidence in perceptual decision making. J Neurosci, 2012. 32(11): p. 3612-28. Drugowitsch, J., R. Moreno-Bote, and A. Pouget, Relation between belief and performance in perceptual decision making. PLoS One, 2014. 9(5): p. e96511. 73. Ma, W.J., et al., Bayesian inference with probabilistic population codes. Nat 74. Neurosci, 2006. 9(11): p. 1432-8. Beck, J.M., et al., Probabilistic population codes for Bayesian decision making. Neuron, 2008. 60(6): p. 1142-52. *75. Ecker, A.S., et al., State dependence of noise correlations in macaque primary visual cortex. Neuron, 2014. 82(1): p. 235-48. Goodfellow, I., Y. Bengio, and A. Courville, Deep Learning. 2016: MIT Press. Bishop, C.M., Pattern recognition and machine learning. 2006, Singapore: Springer. 76. 77. *78. Lin, I.C., et al., The Nature of Shared Cortical Variability. Neuron, 2015. 87(3): p. 80. 81. 79. 644-56. Rabinowitz, N.C., et al., Attention stabilizes the shared gain of V4 populations. Elife, 2015. 4: p. e08998. Softky, W.R. and C. Koch, The highly irregular firing of cortical cells is inconsistent with temporal integration of random EPSPs. J Neurosci, 1993. 13(1): p. 334-50. Tolhurst, D.J., J.A. Movshon, and A.F. Dean, The statistical reliability of signals in single neurons in cat and monkey visual cortex. Vision Res, 1983. 23(8): p. 775-85. Geisler, W.S. and D.G. Albrecht, Visual cortex neurons in monkeys and cats: detection, discrimination, and identification. Vis Neurosci, 1997. 14(5): p. 897-919. *83. Moreno-Bote, R., Poisson-like spiking in circuits with probabilistic synapses. PLoS 82. Comput Biol, 2014. 10(7): p. e1003522. *84. Goris, R.L., J.A. Movshon, and E.P. Simoncelli, Partitioning neuronal variability. 85. 86. 87. Nat Neurosci, 2014. 17(6): p. 858-65. Cohen, M.R. and A. Kohn, Measuring and interpreting neuronal correlations. Nat Neurosci, 2011. 14(7): p. 811-9. Kohn, A., et al., Correlations and Neuronal Population Information. Annu Rev Neurosci, 2016. 39: p. 237-56. Kanitscheider, I., et al., Measuring Fisher information accurately in correlated neural populations. PLoS Comput Biol, 2015. 11(6): p. e1004218. 88. Yatsenko, D., et al., Improved estimation and interpretation of correlations in neural circuits. PLoS Comput Biol, 2015. 11(3): p. e1004083. *89. Rigotti, M., et al., The importance of mixed selectivity in complex cognitive tasks. *90. 91. 92. 93. 94. 95. 96. 97. Nature, 2013. 497(7451): p. 585-90. Ince, R.A., S. Panzeri, and C. Kayser, Neural codes formed by small and temporally precise populations in auditory cortex. J Neurosci, 2013. 33(46): p. 18277-87. Panzeri, S., et al., Neural population coding: combining insights from microscopic and mass signals. Trends Cogn Sci, 2015. 19(3): p. 162-72. Latimer, K.W., et al., Single-trial spike trains in parietal cortex reveal discrete steps during decision-making. Science, 2015. 349(6244): p. 184-7. Latimer, K.W., et al., Response to Comment on "Single-trial spike trains in parietal cortex reveal discrete steps during decision-making". Science, 2016. 351(6280): p. 1406. Shadlen, M.N., et al., Comment on "Single-trial spike trains in parietal cortex reveal discrete steps during decision-making". Science, 2016. 351(6280): p. 1406. Bollimunta, A., D. Totten, and J. Ditterich, Neural dynamics of choice: single-trial analysis of decision-related activity in parietal cortex. J Neurosci, 2012. 32(37): p. 12684-701. Cunningham, J.P. and B.M. Yu, Dimensionality reduction for large-scale neural recordings. Nat Neurosci, 2014. 17(11): p. 1500-9. Kobak, D., et al., Demixed principal component analysis of neural population data. Elife, 2016. 5. *98. Boerlin, M., C.K. Machens, and S. Deneve, Predictive coding of dynamical variables in balanced spiking networks. PLoS Comput Biol, 2013. 9(11): p. e1003258. 99. Moreno-Bote, R. and J. Drugowitsch, Causal Inference and Explaining Away in a Spiking Network. Sci Rep, 2015. 5: p. 17531. 100. Grabska-Barwinska, A., et al., A probabilistic approach to demixing odors. Nat Neurosci, 2016. 20(1): p. 98-106.
1906.01767
1
1906
2019-06-05T00:41:06
Codes, communication and cognition
[ "q-bio.NC" ]
Brette (2019) criticizes the notion of neural coding because it seems to entail that neural signals need to be decoded by or for some receiver in the head. If that were so, then neural coding would indeed be homuncular (Brette calls it dualistic), requiring an entity to decipher the code. But I think the plea of Brett to think instead in terms of complex, interactive causal throughput is preaching to the converted. Turing (not Shannon) has already shown the way. In any case, the metaphor of neural coding has little to do with the symbol grounding problem.
q-bio.NC
q-bio
Commentary on "Is coding a relevant metaphor for the brain?" (Brette 2019). To appear in Behavioral and Brain Sciences. Codes, communication and cognition Stevan Harnad Department of Psychology Université du Québec à Montréal and Department of Electronics and Computer Science University of Southamton [email protected] Abstract: Brette criticizes the notion of neural coding because it seems to entail that neural signals need to be "decoded" by or for some receiver in the head. If that were so, then neural coding would indeed be homuncular (Brette calls it "dualistic"), requiring an entity to decipher the code. But I think Brette's plea to think instead in terms of complex, interactive causal throughput is preaching to the converted. Turing (not Shannon) has already shown the way. In any case, the metaphor of neural coding has little to do with the symbol grounding problem. Both Shannon's (1948) information and Turing's (1936) computation are important in cognitive science. Shannon is concerned with the faithfulness of signal transmission in communication and Turing is concerned with what algorithms can do. Cognitive science is concerned with what organisms (hence their brains) can do, and how. Cells (including neurons) transmit signals. This is already true in plants (Baluska & Mancuso 2009) and of course also in machines. And organisms certainly do things. Which of the things organisms do are "cognitive" and which are "vegetative" is mostly just a definitional matter, but it is probably overstretching the notion to say that paramecia or hearts are "cognizing." The examples are nevertheless instructive for cognitive science, because paramecia, hearts and organisms with brains are all systems that can do things. So are computers and robots, for that matter. Hence finding a causal explanation of how one of them does what it does may provide useful lessons for explaining the others. Let's start with the heart, an example used by Brette. What does the heart do? It pumps blood. No metaphors. The heart literally pumps blood, and cardiac science has successfully reverse- engineered the heart (to a close approximation). We know how the heart does it -- and part of the proof that we know how is that we can apply and test our hypotheses about how the heart pumps blood by building a synthetic model of a heart, plugging it into the heart's inputs and outputs, and testing whether it can pump blood. If it can, the artificial heart passes the "Turing Test" for cardiac function. 1 So what does the (human) brain (and body) pump? Human behavior. Or, rather, human behavioral capacity. What people can do. Let's forget about what portion of that capacity counts as cognitive and what proportion is just vegetative (like cardiac function): It all consists of the capacity of a (living) system to do certain things. Now the challenge is to explain how. Turing (1950) provided the ground rules: You have an explanation if you can design a system that can do everything a human being can do, indistinguishably -- to a human -- from a human. If your interest is just in "cognitive" capacities, then just generate those, ignoring the vegetative capacities (or at least those that are not essential for generating the cognitive capacities). Cognition, like Justice Potter Stewart's pornography, may be hard to define, but we know it when we see it. And the capacity to interact with the dynamic world of objects and events and their properties (including words describing those objects, events and properties) indistinguishably from the way humans do, is surely cognitive, if anything is. There is one more thing: Humans don't just do: They also feel. It feels like something, to a human, to be seeing and doing what humans can see and do. But the capacity to feel eludes Turing's program for cognitive science. It's something our brains pump invisibly. Turing (1950) accordingly brackets it. But it keeps making disruptive peekaboo appearances in our attempts to reverse-engineer cognition, as we shall see. One of the main hypotheses about how the brain pumps cognitive capacity is via computation, Turing computation. Computation is the manipulation of "symbols" (arbitrary formal objects) on the basis of rules operating only on the symbols' shapes ("syntax"), not their meanings ("semantics"), in order to generate certain symbolic outputs from certain symbolic inputs. That's what algorithms do. (An intuitive example is the rule we all learned in school for extracting the roots of quadratic questions: "minus b plus or minus the square root of…".) Algorithms are like recipes: apply them to the symbolic ingredients and you can explain how to bake a symbolic cake. Computation is very powerful; just about everything in the universe can be encoded symbolically and explained computationally, including cardiac function. The right algorithm can pump symbolic blood. And you can show that the algorithm really works by applying it to build a synthetic heart that really passes the cardiac Turing Test and pumps blood. But to do that, you have to "interpret" the symbolic code and implement it in material form, just as a formal recipe for a cake needs to be implemented in material form, using the real ingredients referred to by the symbols, in order to generate a real cake. So, despite its enormous power, computation cannot be all there is to cognition. Searle (1980) showed, famously (in this journal), that a computer is not cognizing even if it can pass the Turing Test (TT) because Searle too could pass the Chinese TT by executing the symbolic code without understanding a word of Chinese. Why can't he understand? Because there is no connection between the symbols in the code and the objects in the world that they are interpretable as being about. Interpretable by whom? The user or the executor of the code. But the meaning itself is not in the code. 2 That is the symbol grounding problem (Harnad 2006). Simple solution: The TT must not be merely symbolic (verbal). It must test not only what the candidate can say, but also all the other things a human cognizer can do in interacting with the objects in the world that the verbal TT is merely chatting about. The candidate has to be a robot. And a Turing robot is not just a computer, manipulating formal symbols; it is a dynamical system, able to interact with the objects in the world. Its symbols are grounded in its capacity to identify and interact with their referents indistinguishably from the way we do. Now to neural "codes": Brette is right that it would be homuncular (although he calls it "dualistic") to think of input to sensory receptors, activity along sensory pathways to sensory and sensorimotor regions in the brain -- and then onward to motor regions and pathways to motor effectors -- as encoded signals being transmitted in order to be decoded by a receiver, as in telegraphic communication of morse code from a sender to a receiver. There is no homunculus on the receiving end. It's all just a dynamic causal process constituting the organism's capacity to do what it can do, some of it output in response to immediate sensory input, some of it generated by endogenous processes. But it is harmless to call the neural activity along sensory input pathways a "neural code." Shannon's communication theory is about the end-to-end fidelity of signal transmission (of analog or digital signals); it is not about cryptography, let alone about the interpretation of computational algorithms or of natural language. To show that there is a substantive issue involved here, Brette would have to show that there is a nontrivial chunk of performance capacity (even the detection of interaural time difference: ITDs) that cannot be explained causally if we insist on calling the activity occurring along the sensory input pathways a "neural code." (Brette's preferred notion of "neural representations," by the way, sounds just as homuncular to me as the idea of neural codes: "representation of what, to whom?" Ditto for "internal model.") Let me close with Brette's fleeting mention of "percepts." This is an instance of the "peekaboo" influence of homuncular thinking. Psychophysics, too, can only study what the organism does (input/output), not whether or how it feels like something to do it. Sensorimotor activity is only perceptual if it is felt. I don't doubt that it feels like something to detect an IDT, just as it feels like something to understand Chinese. But although symbol-grounding and Turing-testing may be "easy" (in principle, if not in practice), explaining how and why organisms feel rather than just do is and remains notoriously hard. References Baluška, F., & Mancuso, S. (2009). Plant neurobiology: from sensory biology, via plant communication, to social plant behavior. Cognitive Processing, 10(1), 3-7. Brette R. (2019) Is coding a relevant metaphor for the brain? Behavioral and Brain Sciences. February 2019 1 -- 44. Harnad, S. (2006). The symbol grounding problem. Encyclopedia of Cognitive Science. Wiley 3 Harnad, S. (2009) The Annotation Game: On Turing (1950) on Computing, Machinery, and Intelligence. In: Epstein, R., Roberts, G., & Beber, G. (Eds.). Parsing the Turing test. Springer Netherlands. Shannon CE (1948) A mathematical theory of communication. Bell system technical journal, 27(3): 379-423. Turing, A. M. (1936). On computable numbers, with an application to the Entscheidungsproblem. Proceedings of the London mathematical society, 2(1), 230-265. Turing , A. M. (1950) Computing Machinery and Intelligence. Mind 49: 433-460. 4
1211.0249
3
1211
2013-01-21T15:27:58
Finite element analysis of neuronal electric fields: the effect of heterogeneous resistivity
[ "q-bio.NC" ]
Simulation of extracellular fields is one of the substantial methods used in the area of computational neuroscience. Its most common usage is validation of experimental methods as EEG and extracellular spike recordings or modeling of physiological phenomena which can not be easily determined empirically. Continuous experimental work has been re-raising the importance of polarization effects between neuronal structures to neuronal communication. As this effects relies on very small potential changes, better modeling methods are necessary to quantify the weak electrical fields in the microscopic scale in a more realistic way. An important factor of influence on local field effects in the hippocampal formation is the heterogeneous resistivity of extracellular tissue. The vast majority of modeling studies consider the extracellular space to be homogeneous while experimentally, it has been shown that the stratum pyramidale has two times higher resistivity then other hippocampal layers. Common simulation methods for extracellular electrical fields based on the point source approximation are bound to describe the resistance of the space with a single, linear factor. We propose that models should be based on the space- and time-dependent Maxwell equations in order to account for heterogeneous properties of the extracellular space and specific arrangements of neurons in dense hippocampal layers. To demonstrate the influence of heterogeneous extracellular resistivity and neuronal spatial orientation on modeling results, we combine solutions of classical compartment models with spatiotemporal PDEs solved by the FEM. With the help of these methods, we show that the inclusion of heterogeneous resistivity has a substantial impact on voltages in close proximity to emitting neurons, increasing the extracellular potentials substantially compared to the homogeneous variant.
q-bio.NC
q-bio
Finite element analysis of neuronal electric fields: the effect of heterogeneous resistivity Pavol Bauer, Sanja Mikulovic, Stefan Engblom, Katarina E Le ao, Frank Rattay, Richardson N Le ao 1 3 1 0 2 n a J 1 2 ] C N . o i b - q [ 3 v 9 4 2 0 . 1 1 2 1 : v i X r a Abstract —Simulation of extracellular fields is one of the sub- stantial methods used in the area of computational neuroscience. Its most common usage is validation of experimental methods as EEG and extracellular spike recordings or modeling of physiological phenomena which can not be easily determined empirically. Continuous experimental work has been re-raising the importance of polarization effects between neuronal struc- tures to neuronal communication. As this effects rely on very small potential changes, better modeling methods are necessary to quantify the weak electrical fields in the microscopic sca le in a more realistic way. An important factor of in fluence on local field effects in the hippocampal formation is the heterogeneous resistivity of extra- cellular tissue. The vast majority of modeling studies consider the extracellular space to be homogeneous while experimentally, it has been shown that the stratum pyramidale has two times higher resistivity than other hippocampal layers. Common simulation methods for extracellular electrical fields based on the poi nt source approximation are bound to describe the resistance of the space with a single scalar. We propose that models should be based on the space- and time-dependent Maxwell equations (Partial Differential Equations, PDEs) in order to account for heterogeneous properties of the extracellular space and speci fic arrangements of neurons in dense hippocampal layers. To demonstrate the in fluence of heterogeneous extracellula r resistivity and neuronal spatial orientation on modeling results, we combine solutions of classical compartment models with spatiotemporal PDEs solved by the Finite Element Method (FEM). With the help of these methods, we show that the inclusion of heterogeneous resistivity has a substantial impact on voltages in close proximity to emitting hippocampal neurons, substantially increasing the change in extracellular potentials compared to the homogeneous variant. Index Terms—extracellular fields finite element method neu- ronal arrangement I . INTRODUCT ION Numerous computational studies have investigated time- varying currents in homogeneous extracellular space [1], [2], [3] as well as the role of neuronal morphology in uniform electric field stimulation [4], [5], [6]. However, most stud ies use models of the extracellular milieu that may not be ac- curately applied when modeling brain structures with highly diverse extracellular resistivity and neuronal arrangement. As emphasized by Lopez-Aguado and Bokil [7], [8] it has been traditionally neglected that currents propagate in all directions in an extracellular medium and that inward and outward currents originate from tissue regions having large re- sistivity differences. A usual argument for this approach is that resistivity in fluences extracellular electrical fields min imally and that extracellular space can be assumed as homogeneous in current-source density (CSD). However, extracellular non- homogeneous resistivity has been shown experimentally in several regions of the brain, for example, in the hippocampus and the cerebellum [9]. Additionally, tissue swelling has been observed after intense neural activity that in turn could lead to an increase in extracellular resistivity [10]. Previous studies [1], [11], [12], [13] have attempted to examine how extracellular electrical fields affects neuronal activity although with the help of quasi-static approximation and an assumed homogeneous tissue resistivity. To quantify the effect of extracellular heterogeneous resis- tivity and neuronal spatial distribution on strength of neu- ral fields, we are proposing a simple modeling pathway to couple compartment-based neural models with the COMSOL Multiphysics simulation environment. Here, we solve time- dependent Maxwell’s equations using the FEM to analyze the change of electrical fields as it occurs in the extracellular space surrounding neurons. Our results indicate that inhomogeneous resistivity of the extracellular milieu significantly in fluences the change of ex- tracellular potentials (EPs) in the hippocampus. By computing the resulting voltage change due to outgoing transmembrane currents of an exemplar CA1 pyramidal cell model in ho- mogeneous and inhomogeneous extracellular resistivity, we observed a maximal difference in EP change of 60% in the hippocampal pyramidal layer. Furthermore, the here proposed method offers the possibility to efficiently simulate the ef fects of superimposed extracellular potentials created by neurons in divergent positions relative to each other. We will also discuss advantages and drawbacks of the used method and propose alternatives and possible improvements towards more realistic modeling of electrical fields of the brain. Pavol Bauer and Stefan Engblom are with Division of Scientifi c Comput- ing, Department of Information Technology, Uppsala University, SE-751 05 Uppsala, Sweden, e-mail: [email protected] Sanja Mikulovic, Katarina E Le ao and Richardson N Le ao are with Department of Neuroscience, Uppsala University Box 593, 751 24 Uppsala, Sweden, e-mail: [email protected] Sanja Mikulovic and Frank Rattay are with Institute for Analysis and Scientific Computing, Vienna University of Technology Vien na, Austria Katarina E Le ao and Richardson N Le ao are with Brain Instit ute, Federal University of Rio Grande de Norte Natal-RN, Brazil I I . MAT ER IAL S AND M E THOD S A. Pyramidal neuron model To demonstrate the effect of trans-membrane currents on the effect of extracellular fields, a Hodgkin-Huxley (HH)-l ike hippocampal CA1 pyramidal cells was adapted from [14] with the addition of a current Ih from [15] and a current Im from [16]. Ih current: u∞ = [1 + exp((V + 76)/7)]−1 , τu = 104/[237 exp((V + 50)/12)+ 17 exp(−(V + 50)/25)] + 0.6, Ih (t, V ) = ghu(t, V ) (V − Eh ). Im current: s∞ = [1 + exp(−(V + 22.53)/10)]−1, τs = 4135.7/[164.64 exp((V − 0.05) · 0.12)+ 0.33 exp(−(V − 0.05)/10)] + 35.66, Im (t, V ) = gm s(t, V ) (V − EK ). (1) (2) (3) (4) (5) (6) Our model contained one somatic (r = 10 µm), 20 dendritic (r = 1 –3 µm, l = 5 µm), two axon initial segments, AI S1 (r = 0.5 µm, l = 60 µm ), AI S2 (r = 0.5 µm, l = 60 µm), and finally 32 axonal compartments (r = 0.5 µm, l = 5 µm), see Figure 1. Note that here AI S1 refers to the part of AIS from 0 to 60 µm, and AI S2 from 60 to 120 µm from the cell body. Additionally, AI S2 compartments contain high N a+ channel density [17]. 20 dendritic compartments are used for the dendritic tree, while additional compartments are included in the branching analysis. The N a+ channel density was varied between 986 and 2943 µA cm2 (corresponding to AI S2 ). The following conductance values were used: gN a = 8, gkdr = 5, gm = 1 mS/cm2 for the soma; gC a = 10, gKC = 15, gKAHP = 0.8, gO = 0.625, gN a = 0.07, gh = 0.2 mS/cm2 for the dendrite; gN a = 50, gkdr = 10, gm = 10 mS/cm2 for the AI S2 and gN a = 9, gkdr = 10, gm = 10 mS/cm2 for the rest of the axon compartments. The leakage conductance was set to 0.1 for all of the compartments. Finally, the equilibrium potentials were set to EN a = 60 mV, EK = -85 mV, Eh = -43 mV and Eleakage = -65 mV. B. Creation of morphology and model coupling The three-dimensional neuronal geometry was constructed in COMSOL Multiphysics 4.3 with the help of the interface to MATLAB ( “LiveLink ”) by morphological additions and boolean unification of simple geometric volumes. As we aim to represent the 3D-morphology as an exact counterpart of the compartmental model, each section is recreated as a cylinder with the same length and diameter as in the compartmental de finition. If the cylinders were added on top of each other with no change of the rotation vector, as shown in Figure 1, no joining geometrical primitives were added in between them. If the rotation differs, as for example in Figure 4, a sphere is added in between the cylinders, followed by a removal of the interior boundaries of both the cylinders and the sphere. The reason for constructing the geometry in this way is that the mesh engine otherwise respects the internal boundaries such that the resulting mesh becomes unnecessary complex. In some cases the occurrence of multiple internal boundaries can even make the meshing procedure abort. Ionic currents of single neuronal compartments model were determined by solving Hodgkin-Huxley Ordinary Differential Equations (HH ODEs) specified in equations 1 to 6 using a 2 2 0 µ m COMPARTMENT MODEL IONIC CURRENT FEM MODEL ICa, IK, IK_AHP, IK_O, INa, Ih ICa, IK, IK_AHP, IK_O, INa, Ih IKdr, Im, Ina IKdr,Kdr IKdr, Im, Im, Ina Ina IKdr, Im, Ina IKdr, Im, Ina Dend 2,...,n DendDend 1 Soma AIS AIS 1 1 AIS 2 Axon 1,...,n Fig. 1. Mapping trans-membrane currents from compartmental model to FEM boundary mesh. A compartmental model containing 20 dendritic, one soma, two axon initial segments and 32 axonal compartments. Ionic currents for each compartment are shown on the left. The compartmental model outputs I (t) which is used as current boundary sources Qj (t) in the corresponding three-dimensional volumes in the FEM model. Runge-Kutta algorithm from the MATLAB ODE-suite (Math- works) with a constant step-size T . The sum of all currents per compartment for each time step of the simulation I0 is afterwards normalized to an absolute value of units A/m2 and stored in a matrix I of size Ncompartments × Ntimesteps . Next, by using the interface to COMSOL, each row of the matrix I was mapped adequately to the corresponding cylindrical domain as a boundary current source Qj (t). Note that we hereby assume that transmembrane currents are the only cause of change of extracellular potential, which is surely not the case in a real neuron, as for example synaptic calcium- mediated currents are suspected to contribute to a large fraction of the extracellular signature [18]. The extracellular volume was modeled by cylindrical ob- jects covering the neuronal morphology with either homo- geneous (0.3 S/m) or heterogeneous resistivity as shown in Figure 2A. In this case we constructed the extracellular volume by taking the union of cylindrical objects with increasing resistivity in the y -axis, according to [7]. It is assumed that the conductivity of extracellular tissue is frequency-independent in the used range of neural activity (10 –100 Hz) [19]. C. Electrostatic formulation and the Finite Element Method To simulate non-homogeneous distributions of electrical fields produced by single neurons and neuronal networks, we used an electrostatic formulation of Maxwell’s equations discretized with finite elements. Chie fly, in the finite eleme nt approach, Maxwell’s equations are solved by discretizing the incorporated volumes (in this case the neuronal compartments) into finite tetrahedral volume elements [20]. We seek the elec- tric field intensity E in terms of the electric scalar potential V , E = −∇V . (7) The relevant dynamic form of the continuity equation with current sources Qj is given by ∂ ρ ∂ t with J and ρ the current density and electric charge density, respectively. Further constitutive relations include ∇ · J = − + Qj , (8) and Ohm’s law D = ε0 εrE, J = σE, (9) (10) in which D denotes the electric flux density . Finally, Gauss’ law states that ∇ · D = ρ. (11) Upon taking the divergence of (10) and using the continuity equation (8) we get + Qj . ∇ · σE = − ∂ ρ ∂ t Rewriting the electric charge density using Gauss’ law together with the constitutive relation (9) and finally applying the g auge condition (7) twice we arrive at the time-dependent potential formulation (12) (13) ∇V (cid:19) = Qj . − ∇ · (cid:18)σ∇V + ε0 εr ∂ ∂ t This is the formulation used in COMSOL Multiphysics [21]. The values for the electric conductance σ and the relative permittivity εr were obtained from [2]. The source currents Qj were computed from the compartmental model as described above. The formulation (13) is efficiently solved by COMSOL’s “Time discrete solver ”, which is based on the observation that the variable W := ∆V satisfies a simple ODE. Solving for W in an independent manner up to time t, it is then straightforward to solve a single static PDE to arrive at the potential V itself. As for boundary conditions we took homogeneous Neu- mann conditions (electric isolation) everywhere except for in a single point which we choosed to be ground (V = 0). In all our simulations this point was placed at the axis of rotation of the enclosing cylindrical extracellular space, and underneath the neuronal geometry. This procedure ensures that the formulation has a unique solution (it is otherwise only specified up to a constant). 3 A tetrahedral mesh was applied to discretize space (using the “ finest” mesh setting; resolution of curvature 0.2, reso - lution of narrow regions 0.8). The simulations were verified against coarser mesh settings in order to ensure a practically converged solution. As a final note, in the Time discrete solv er the time step was set to the same step size T as used in the ODE-based solution of the Hodgkin-Huxley equations, thus ensuring a correct transition between both simulation environments. I I I . RE SULT S A. The effect of heterogeneous extracellular fields In order to examine the in fluence of heterogeneous ex- tracellular space, we first constructed a three-dimensiona l active neuron model in an homogeneous and heterogeneous extracellular milieu (Figure 2). Neurons in the hippocampus have an intricate spatial orientation that propitiates strong field potentials: high density of pyramidal cell dendrites running in parallel in the stratum radiatum (SR), densely packed pyramidal cell somas in the stratum pyramidale (SP) while pyramidal cell axons run almost in parallel or crossing each other in SP or stratum oriens (SO). We measured the voltage on four de fined point probes placed parallel to the dendrites, soma, AI S2 and axon terminal compartments during the peak of an action potential (AP) of 80mV, by varying the distance between active neuron and point probes from 1 µm to 80 µm (Figure 2A). In the first set of simulations, we analyzed the effect of the aforementioned four neuronal regions on the de fined poin t probes assuming a widely accepted homogeneous resistivity of 350 Ω cm [2], [3]. Note that here applied boundary currents of three-dimensional neuronal compartments correspond to the peak transmembrane current during an AP. By doing so, the peak voltage of 0.25 mV was obtained in the point probe parallel to the AI S2 compartment, followed by soma, dendrite and axon terminal. Consequently, the heterogeneous case was examined by placing the neuron and point probes into a heterogeneous extracellular space representing hippocampal spatial order, where resistivity values for different strata were obtained from [7] and shown in Figure 2A. The active neuron was positioned in the center of SP and point probes were moved along the x axis of the extracellular space. In the case of non-homogeneous extracellular resistivity (Figure 2C) the largest voltage change was measured in point probes parallel to AI S2 , although the values in point probes placed in SP were 60% larger in close distance than in the homogeneous extracellular space scenario and 28% higher considering the spatial mean over the total distance. The point probes placed in SO and SR were as well affected by the higher restivity of the pyramidal layer, showing an average increase of 4% in parallel to the axon and 7% in parallel to the dendrite. The appreciable difference between the voltage changes suggests that non- homogeneous distribution of resistivity is an important aspect of extracellular field effects. A B D S AIS 2 A 0.2 0.15 0.1 0.05 ) V m ( e g a t l o V SR SP SO 100 300 200 500 400 Resistivity (Ωcm) 600 D S 0.4 0.3 0.2 0.1 0 0 0.2 0.15 0.1 0.05 0 0 10 20 30 40 A 10 30 20 Distance (µ m) 40 0 0 10 20 30 40 AIS 2 ) V m ( e g a t l o V 0.4 0.3 0.2 0.1 0 0 10 30 20 Distance (µm) 40 Fig. 2. Heterogeneous extracellular resistivity modulates the strength of polarization effects in proximity of the firing neuron . A, Schematic representation of a model neuron impact on defined point prob es. Point probes act as receivers and are being shifted away from the active neuron along the x-axis (left). Non-homogeneous resistivity distribution of hippocampal regions (right). B, Relationship between membrane potential at different point probes affected by dendritic (D), somatic (S), axon initial segment (AI S2 ) and axon terminal (A) neuronal compartments (color coded as in A) for distance varying between 0 and 80 µm. Dashed: the resulting voltage change assuming homogenous extracellular resistivity (= 350 Ω cm), while solid represents the result of assuming a heterogenous resistivity. B. Analysis in respect to spatial orientation and neural mor- phology According to the superposition principle electrical fields (EFs) are produced by summation of single neuronal activity. Thus, interaction between neuron and field is the mutual strongly modulated by the spatial orientation of neuronal assemblies [22]. To analyze EFs considering a realistic spatial order, we began simulating two neurons with parallel axons (Figure 3A) where N eurona represents an active structure with boundary current source Qj mapped to its surface while N euronp is a passive measurement structure. We measured Ve at four points along the axon of N euronp and detected the maximal voltage amplitude of 0.28 mV. Note that here Ve corresponds to extracellular voltage measured on the cell membrane. Next we tested the in fluence of four neighboring cells with parallel axons on the N euronp (Figure 3B). In this case, having four active neuronal neighbors (distance between the axons =10 µm) firing non-synchronously, a maximal voltage of 0.49 mV in homogeneous and 0.68 mV in heteroge- neous extracellular medium of N euronp was computed. Syn- chronous firing of adjacent neurons produced a peak voltage of 0.62 mV in homogeneous and 0.84 mV in heterogeneous 0.3 0 −0.1 0.3 0 −0.1 0 H o m o g e n o u s H e t e r o g e n o u s Time (ms) 100 Unsynchronized Synchronized 0.7 0 -0.45 0.85 0 -0.6 40 0 Time (ms) N 1 N 2 N 3 N 4 N 1 N 2 N 3 N 4 40 Unsynchronized Synchronized 1.2 0 -0.8 1.4 0 -0.9 40 N 1 N 2 N 3 N 4 40 0 Time (ms) N 1 N 2 N 3 N 4 40 4 H o m o g e n o u s H e t e r o g e n o u s H o m o g e n o u s H e t e r o g e n o u s Neuron p Neuron p Neuron p 0.5 0 -0.3 0.7 0 -0.35 0 0.8 0 -0.5 0.9 0 -0.6 0 A SO SP SR B SO SP SR C SO SP SR Fig. 3. Analysis of extracellular fields in dependence of spike timi ng and spatial orientation. A. Potential change calculated at different regions (colored circles) of a passive neuron (N euronp ) in response to EPs triggered from action potentials in the initial phase of an active neuron. Lines in the right panel are color coded as the circles in the left panel. Somas were placed at the stratum pyramidale and extracellular resistivity followed the same distribution as shown in Figure 2C. Upper trace of the right panel shows potential changes in homogeneous, while lower trace in heterogeneous extracellular space. B, Same as in A but the N euronp surrounded by four active neurons. Potential change at different regions of N euronp when four active neurons fire asynchronously (middle panel) and synch ronously (right panel). Insets on the middle and right panels show the firing of active neuron s). C, Same as in B, but the active neurons where oblique (but not intersecting, nearest interaxonal distance = 2 µm) at the position of AI S2 to N euronp . medium of AI S2 . When axons crossed each other at the AI S2 region (Figure 3C), we observed a maximal voltage amplitude of 0.89 mV in the AI S2 of Axonp during asynchronous activity of the neighboring cells, whereas synchronous activity produced 1.41 mV in heterogeneous medium. These results suggest that for analyzed neuronal arrangement soma and AI S2 have considerable in fluence on the strength of neural fields for small distances ( ≤ 10 µm) and that values in heterogeneous extracellular space are in mean ∼ 25% larger than in homogeneous extracellular medium when averaged over time. Furthermore, we simulated the in fluence of pyramidal neuron position within different hippocampal layers on the strength of neural fields (Figure 4). We positioned somas in SP with the smallest distance between the two axon initial segments con fined to SO. This geometrical arrangement was motivated by confocal images of two proximal pyramidal neurons located parallel to each other with an axonal bend A 27.43 µm 2 µm 13.53 µm 6.99 µm B A SR SP SO C SR SP SO 0 0.1 0.2 0.3 0.4 0.5 0 0.1 0.2 0.3 0.4 0.5 Voltage (mV ) Voltage (mV ) Fig. 4. Modeling the effect of bended axons. A, Confocal image of two neurons in the hippocampus filled with neurobiotin for post hoc spatial analysis showing two parallel neurons with bent axons situated at a minimal distance of 2 µm. B, A simplified representation of both neurons when the nearest interaxonal region was confined in stratum oriens. I n this case, an action potential in the active neuron generates a voltage transient of 0.25 mV at the passive neuron soma and 0.35 mV at the neuron AI S2 . The figure shows the extracellular voltage distribution at the peak amplitude of the active cell action potential. C, Same as B, but axon initial segments were confined at the stratum pyramidale (higher resistivity). In this case, a 0.25 mV and 0.45 mV voltage transients were observed at the passive neuron soma and axon, respectively. starting about 50 µm behind the soma (Figure 4A). As known from previous analytical studies [23] the bend of axonal structures generally amplifies endogenous fields aro und neurons. We were able to con firm this effect in our simulation as we simulated the field effect resulting from trans-membra ne currents of an active neuron measured on the parallel neuron. In this scenario, peak voltage of 0.35 mV was calculated at the position of AI S2 while the smallest voltages were registered at the passive neurons between the dendrites (0.12 mV) in SR and axon terminals (0.16 mV). If AI S2 were con fined to SP, the potential computed at the passive neuron was 25% larger than in the case in which AI S2 were positioned in the SO (Figure 4B). IV. D I SCU S S ION In this work, we use the Finite Element Method coupled to the HH-equations to simulate how neuronal geometry, arrangement and heterogeneous extracellular properties affect the strengths of neural fields. We first show that there is a difference in voltage transients produced by firing of neigh - boring cells if the extracellular space is considered to be homogeneous or heterogeneous. Additionally, we demonstrate that the spatial orientation of specific cellular compartme nts is an important determinant of the strength of neural fields. 5 In our computations, the highest change of extracellular potential arises in the pyramidal layer, in proximity to the AI S2 compartment. Action potentials are generally initiated in the AIS due to a higher density of voltage-gated N a+ channels [17], [24], [25], which is re flected by the parameter setting s for this compartment in the occupied neuronal model. Ad- ditionally, the spatial arrangement of hippocampal pyramidal neurons also propitiates the proximity of AI S2 in both SP and SO (see Figure 4, [26]). Holt and Koch [27] showed that interactions near cell bodies are more important than interactions between axons by using standard one-dimensional cable theory and volume conductor theory. Another study [28] reported a 4.5 mV change (in the AIS) caused by extracellular interactions, a change more than 4× larger than what we have found in our simulations of heterogeneous extracellular media. Additionally, by applying analytical methods, Bokil et al [8] have shown that, in the olfactory system, an action potential of 100 mV amplitude in one axon could produce depolarization in other axons in the bundle sufficient to initiate an action potential. Howev er, using FEM simulations it was not possible to trigger spikes in neurons solely by electrical fields mediated by in- and outflo w of trans-membrane currents, in agreement with the work by Traub and colleagues (in which maximal voltages in a sink axon during synchronized activity of four neighboring neurons is ≈ 1.2mV ). Hence, simulations relying on the point source approxima- tion to describe hippocampal neural fields may be distorted a s the potential change is more than 28% greater in stratum pyra- midale and in average 6% greater in other hippocampal layers if a heterogeneous tissue is used instead of a homogeneous one [7], [1]. For varying extracellular resistivity a numerical procedure referred to as ‘the method of images’ [29] has been proposed as an extension to the point source equation. Although this method has been used in computations of extracellular action potentials in a previous study [11], its practical use is limited to a rather low number of resistivity layers and non-complex geometry. The requirement of FEM to model passive current flow between neurons was suggested elsewhere [2], [3], but im- plementation issues and the lack of adequate software tools may have precluded its usage in the past. Extracellular resistivity could also contribute to the ampli- tude of local field potentials (LFP), and in fact, in vitro LFP registered in hippocampal slices are greater in the SP than in other layers [30]. Interestingly, we observed that hippocampal slices in interface-type recording chambers (where slices are not completely submerged in the buffer solution [31], [32]) are more than 60% less conductive than in the chambers where slices are submerged (unpublished results). This could help explaining why LFP recorded in interface-type cambers are far greater than in submerged-type counter parts [16], [32]. Here we show that the use of FEM software (COMSOL Multyphysics), with an interface to Hodgkin and Huxley- type ODE model in Matlab, is a powerful tool to verify macroscopic effects of neural fields. However, this approac h may fail to simulate small nuances of extracellular interactions (e.g. ion channels apposing active compartments may suffer more from the effect of passive current flow than channels in the opposite side). Nonetheless, simulations using PDEs could increase the level of realistic models of membrane dynamics. The idea of translating membrane dynamics to PDE was proposed by Hodgkin and Huxley themselves [33], and later pursued by Kashef and Bellman [34]. However, implementing membrane dynamics in combination with the Maxwell equation interface of the PDE simulator has been, to our experience, quite cumbersome. For example, due to the relatively high computational cost of the solution phase of the finite element method, a representation of fully reconstruc ted morphologies with volumetric cylindrical elements was not possible. The high amount of (small) cylindrical elements required to accurately model complex dendritic branches typ- ically caused the meshing algorithm to break. The software used in our simulations (COMSOL Multi- physics) has helped to popularize the use of FEM in neurosci- entific problems [35], [36]. However, the software is geared towards industrial applications and the steep learning curve associated to the adaptation of COMSOL to neuroscience related problems may preclude its widespread usage by the neuroscience community. Possible combination of Neuron Simulation Environment (NSE) with a FEM simulator would offer great possibilities by allowing both realistic neural mor- phologies and extracellular space modeling. However, cur- rently the existence of a direct interface between NSE and a FEM simulator has not been reported; future studies should focus on this task. Furthermore, hippocampal extracellular stimulation has been proposed as therapeutic strategy to con- trol several disorders, including temporal lobe epilepsy [37], [38]. Developing models that consider as many as possible physical tissue properties would help to improve extracellular stimulation in epilepsy patients. In summary, our work adds to the recently published studies attempting to reveal important parameters determining the strength of extracellular electric fields. Models that do not use space and time-dependent differential equations when modeling neuronal interactions may have failed to replicate the changes in measured voltages caused by passive current flow in heterogeneous extracellul ar tissue, especially when more than two neurons are modeled simultaneously. ACKNOW L EDGMENT This work was supported by Kjell and M arta Beijers Foun- dation and the BMWF Marietta Blau stipend. PB and SE were supported by the Swedish Research Council within the UPMARC Linnaeus center of Excellence. RE F ERENCE S [1] C. C. McIntyre and W. M. Grill, “Excitation of central ner vous system neurons by nonuniform electric fields,” Biophysical Journal, vol. 76, no. 2, pp. 878–888, Feb. 1999, PMID: 9929489. [Online]. Avai lable: http://www.ncbi.nlm.nih.gov/pubmed/9929489 [2] C. B ´edard, H. Kr oger, and A. Destexhe, “Modeling extra cellular field frequency-filtering properties of extra cellular potentials and the no. space,” Journal, vol. 86, 3, pp. 1829– Biophysical 1842, Mar. 2004, PMID: 14990509. [Online]. Available: http://www.ncbi.nlm.nih.gov/pubmed/14990509 6 [3] — in brain “Model of low-pass filtering of local field potentials —, tissue,” Physical Review. E, Statistical, Nonlinear, and Soft Matter Physics, vol. 73, no. 5 Pt 1, p. 051911, May 2006, PMID: 16802971. [Online]. Available: http://www.ncbi.nlm.nih.gov/pubmed/16802971 [4] C. Y. Chan, J. Hounsgaard, and C. Nicholson, “Effects of e lectric fields on transmembrane potential and excitability of turtl e cerebellar The Journal of Physiology, vol. 402, purkinje cells in vitro,” pp. 751–771, Aug. 1988, PMID: 3236254. [Online]. Available : http://www.ncbi.nlm.nih.gov/pubmed/3236254 [5] T. Radman, Y. Su, J. H. An, L. C. Parra, and M. Bikson, “Spik e timing amplifies the effect of electric fields on neurons: imp lications for endogenous field effects,” The Journal of Neuroscience: The the Society for Neuroscience , vol. 27, no. 11, Official Journal of pp. 3030–3036, Mar. 2007, PMID: 17360926. [Online]. Availa ble: http://www.ncbi.nlm.nih.gov/pubmed/17360926 [6] T. Radman, R. L. Ramos, J. C. Brumberg, and M. Bikson, “Rol e of cortical cell type and morphology in subthreshold and suprathreshold uniform electric field stimulation in vitro,” Brain Stimulation, vol. 2, no. 4, pp. 215–228, 228.e1–3, Oct. 2009, PMID: 20161507. [On line]. Available: http://www.ncbi.nlm.nih.gov/pubmed/20161507 [7] L. L ´opez-Aguado, J. M. Ibarz, and O. Herreras, “Activit y-dependent changes of tissue resistivity in the CA1 region in vivo are layer- Neuroscience, vol. 108, specific: modulation of evoked potentials,” no. 2, pp. 249–262, 2001, PMID: 11734358. [Online]. Availab le: http://www.ncbi.nlm.nih.gov/pubmed/11734358 [8] H. Bokil, N. Laaris, K. Blinder, M. Ennis, and A. Keller, “ Ephaptic interactions in the mammalian olfactory system,” The Journal of Neuroscience: The Official Journal of the Society for Neuros cience, vol. 21, no. 20, p. RC173, Oct. 2001, PMID: 11588203. [Online]. Available: http://www.ncbi.nlm.nih.gov/pubmed/11588203 [9] Y. C. Okada, J. C. Huang, M. E. Rice, D. Tranchina, and C. Nicholson, “Origin of the apparent tissue conductivity in the molecula r and granular layers of the in vitro turtle cerebellum and the interpretation of current source-density analysis,” Journal of Neurophysiology, vol. 72, no. 2, pp. 742–753, Aug. 1994, PMID: 7983532. [Online]. Avai lable: http://www.ncbi.nlm.nih.gov/pubmed/7983532 [10] A. M. Autere, K. Lamsa, K. Kaila, and T. Taira, “Synaptic activation of GABAA receptors induces neuronal uptake of ca2+ in adult rat Journal of Neurophysiology, vol. 81, no. 2, hippocampal slices,” pp. 811–816, Feb. 1999, PMID: 10036281. [Online]. Availabl e: http://www.ncbi.nlm.nih.gov/pubmed/10036281 “On [11] C. Gold, D. A. Henze, C. Koch, and G. Buzs ´aki, the the origin of action potential waveform: A extracellular modeling study,” Journal of Neurophysiology, vol. 95, no. 5, pp. 3113–3128, May 2006, PMID: 16467426. [Online]. Availab le: http://www.ncbi.nlm.nih.gov/pubmed/16467426 [12] C. Gold, D. A. Henze, “Using extracellular and C. Koch, to constrain compartmental models,” action potential recordings vol. 23, no. 1, pp. Journal of Computational Neuroscience, 39–58, Aug. 2007, PMID: 17273940. [Online]. Available: http://www.ncbi.nlm.nih.gov/pubmed/17273940 [13] C. A. Anastassiou, S. M. Montgomery, M. Barahona, G. Buzs ´aki, and C. Koch, “The effect of spatially inhomogeneous extrace llular electric fields on neurons,” The Journal of Neuroscience: The the Society for Neuroscience , vol. 30, no. 5, Official Journal of pp. 1925–1936, Feb. 2010, PMID: 20130201. [Online]. Availa ble: http://www.ncbi.nlm.nih.gov/pubmed/20130201 “Intrinsic and network rhyth mogenesis in [14] P. F. Pinsky and J. Rinzel, a reduced traub model for CA3 neurons,” Journal of Computational Neuroscience, vol. 1, no. 1-2, pp. 39–60, Jun. 1994, PMID: 8792224. [Online]. Available: http://www.ncbi.nlm.nih.gov/pubmed/8792224 [15] R. N. Leao, K. Svahn, A. Berntson, and B. Walmsley, “Hyperpolarization-activated (I) currents in auditory br ainstem neurons of normal and congenitally deaf mice,” The European Journal of Neuroscience, vol. 22, no. 1, pp. 147–157, Jul. 2005, PMID: 16029204. [Online]. Available: http://www.ncbi.nlm.nih.gov/pubmed/16029204 [16] R. N. Le ao, H. M. Tan, and A. Fisahn, “Kv7/KCNQ channels control action potential phasing of pyramidal neurons during hippocampal gamma oscillations in vitro,” The Journal of Neuroscience: The the Society for Neuroscience , vol. 29, no. 42, Official Journal of pp. 13 353–13 364, Oct. 2009, PMID: 19846723. [Online]. Avai lable: http://www.ncbi.nlm.nih.gov/pubmed/19846723 [17] L. M. Palmer and G. J. Stuart, i nitiation “Site of action potential in layer 5 pyramidal neurons,” The Journal of Neuroscience: The the Society for Neuroscience , vol. 26, no. 6, Official Journal of pp. 1854–1863, Feb. 2006, PMID: 16467534. [Online]. Availa ble: http://www.ncbi.nlm.nih.gov/pubmed/16467534 7 epilepsy,” Stereotactic and functional neurosurgery, vol. 89, no. 2, pp. 111–122, Apr. 2011, PMID: 21336007. [38] H. Luna-Mungua, A. Meneses, F. Pea-Ortega, A. Gaona, and L. Rocha, “Effects of hippocampal high-frequency electrical stimul ation in mem- ory formation and their association with amino acid tissue content and release in normal rats,” Hippocampus, vol. 22, no. 1, pp. 98–105, Jan. 2012, PMID: 20882549. [18] K. C. Buszaki G, Anastassiou CA, in the CN S “Field effects play functional roles,” Nature Neurosciecne, vol. 6, 2012. [Online]. Available: http://www.ncbi.nlm.nih.gov/pubmed/22595786 “Intrin sic and G. T. Einevoll, [19] H. Lind ´en, K. H. Pettersen, filtering gives dendritic local low-pass power spectra of fie ld Journal of Computational Neuroscience, vol. 29, no. 3, potentials,” pp. 423–444, Dec. 2010, PMID: 20502952. [Online]. Availabl e: http://www.ncbi.nlm.nih.gov/pubmed/20502952 [20] J. Jin, The Finite Element Method in Electromagnetics, 2nd ed. New York: John Wiley & Sons, 2002. [21] AC/DC Module User’s Guide, Comsol, 2012, version 4.3. [22] F. Fr ohlich and D. A. McCormick, fields “Endogenous electric may guide neocortical network activity,” Neuron, vol. 67, no. 1, pp. 129–143, Jul. 2010, PMID: 20624597. [Online]. Availabl e: http://www.ncbi.nlm.nih.gov/pubmed/20624597 “A model for the polariza tion of [23] D. Tranchina and C. Nicholson, neurons by extrinsically applied electric fields,” Biophysical Journal, vol. 50, no. 6, pp. 1139–1156, Dec. 1986, PMID: 3801574. [Onl ine]. Available: http://www.ncbi.nlm.nih.gov/pubmed/3801574 central [24] B. P. Bean, “The action potential in mammalian vol. Nature Reviews. Neuroscience, neurons,” 6, pp. 8, no. 451–465, [Online]. Available: Jun. 2007, PMID: 17514198. http://www.ncbi.nlm.nih.gov/pubmed/17514198 [25] M. H. P. Kole and G. J. Stuart, thres hold “Is action potential Nature Neuroscience, vol. 11, no. 11, lowest in the axon?” pp. 1253–1255, Nov. 2008, PMID: 18836442. [Online]. Availa ble: http://www.ncbi.nlm.nih.gov/pubmed/18836442 [26] S. Ram ´on y Cajal, Histologie du Systeme Nerveux de l’Homme et des Vertebretes, 1911. [27] G. R. Holt and C. Koch, “Electrical interactions via the extracellular Journal of Computational Neuroscience, potential near cell bodies,” vol. 6, no. 2, pp. 169–184, Apr. 1999, PMID: 10333161. [Onlin e]. Available: http://www.ncbi.nlm.nih.gov/pubmed/10333161 [28] R. D. Traub, F. E. Dudek, C. P. Taylor, and W. D. “Simulation of hippocampal afterdischarges sync hronized Knowles, by electrical interactions,” Neuroscience, vol. 14, no. 4, pp. 1033–1038, Apr. [Online]. Available: 2987752. 1985, PMID: http://www.ncbi.nlm.nih.gov/pubmed/2987752 [29] E. Weber, Electro Magnetic Fields: Theory, Applications. New York: John Wiley & Sons, 1950. [30] A. Fisahn, F. G. Pike, E. H. Buhl, and O. Paulsen, “Cholin ergic induction of network oscillations at 40 hz in the hippocampus in vitro,” Nature, vol. 394, no. 6689, pp. 186–189, Jul. 1998, PMID: 9671302. [Online]. Available: http://www.ncbi.nlm.nih.gov/pubmed/9671302 “Volume-conducted epileptiform events Inaba and M. Avoli, [31] Y. adjacent between necortical slices in an interface tissue chamber,” Journal of Neuroscience Methods, vol. 151, no. 2, pp. 287–290, Mar. 2006, PMID: 16143402. [Online]. Availabl e: http://www.ncbi.nlm.nih.gov/pubmed/16143402 [32] N. H ´ajos, T. J. Ellender, R. Zemankovics, E. O. Mann, R. Exley, “Maintaining netwo rk S. J. Cragg, T. F. Freund, and O. Paulsen, activity in submerged hippocampal slices: importance of oxygen supply,” The European Journal of Neuroscience, vol. 29, no. 2, pp. 319–327, [Online]. Availabl e: Jan. 2009, PMID: 19200237. http://www.ncbi.nlm.nih.gov/pubmed/19200237 “A quantitative descript ion [33] A. L. HODGKIN and A. F. HUXLEY, current of membrane and its application to conduction and excitation in nerve,” The Journal of Physiology, vol. 117, no. 4, pp. 500–544, Aug. 1952, PMID: 12991237. [Online]. Availabl e: http://www.ncbi.nlm.nih.gov/pubmed/12991237 [34] B. Kashef and R. Bellman, “Solution of the partial diffe rential equation of the Hodgkin-Huxley model using differential quadrature,” Mathemat- ical Biosciences, vol. 19, no. 1-2, pp. 1–8, 1974. [35] T. C. Zhang and W. M. Grill, “Modeling deep brain stimula tion: the point source approximation versus realistic representation of vol. electrode,” of Neural Engineering, 7, no. 6, Journal p. 066009, Dec. 2010, PMID: 21084730. [Online]. Available: http://www.ncbi.nlm.nih.gov/pubmed/21084730 [36] N. Yousif, N. Purswani, R. Bayford, D. Nandi, P. Bain, and X. Liu, the deep brain stimulation “Evaluating the impact of induced electric field on subthalamic neurons: a computatio nal Journal of Neuroscience Methods, vol. 188, no. 1, modelling study,” pp. 105–112, Apr. 2010, PMID: 20116398. [Online]. Availabl e: http://www.ncbi.nlm.nih.gov/pubmed/20116398 [37] P. H. Stypulkowski, J. E. Giftakis, and T. M. Billstrom, “Development of a large animal model for investigation of deep brain stimulation for
1002.4903
1
1002
2010-02-26T00:33:16
Sparse incomplete representations: A novel role for olfactory granule cells
[ "q-bio.NC" ]
Mitral cells of the olfactory bulb form sparse representations of the odorants and transmit this information to the cortex. The olfactory code carried by the mitral cells is sparser than the inputs that they receive. In this study we analyze the mechanisms and functional significance of sparse olfactory codes. We consider a model of olfactory bulb containing populations of excitatory mitral and inhibitory granule cells. We argue that sparse codes may emerge as a result of self organization in the network leading to the precise balance between mitral cells' excitatory inputs and inhibition provided by the granule cells. We propose a novel role for the olfactory granule cells. We show that these cells can build representations of odorant stimuli that are not fully accurate. Due to the incompleteness in the granule cell representation, the exact excitation-inhibition balance is established only for some mitral cells leading to sparse responses of the mitral cell. Our model suggests a functional significance to the dendrodendritic synapses that mediate interactions between mitral and granule cells. The model accounts for the sparse olfactory code in the steady state and predicts that transient dynamics may be less sparse.
q-bio.NC
q-bio
Sparse incomplete representations: A novel role for olfactory granule cells Alexei A. Koulakov1 and Dmitry Rinberg2 1Cold Spring Harbor Laboratory, Cold Spring Harbor, NY 11724 2HHMI Janelia Farm Research Campus, HHMI, Ashburn, VA 20147 Mitral cells of the olfactory bulb form sparse representations of the odorants and transm it this information to the cortex. The olfactory code carried by the mitral cells is sparser than the inputs that they receive. In this study we analyze the mechanisms and functional significance of sparse olfactory codes. We consider a model of olfactory bulb containing populations of excitatory m itral and inhibitory granule cells. W e argue that sparse codes may emerge as a result of self organization in the network leading to the precise balance between m itral cells' excitatory inputs and inhibition prov ided by the granule cells. We propose a novel role for the olfactory granule cells. We show that these cells can build representations of odorant stimuli that are not fully accurate. Due to the incompleteness in the granule cell representation, the exact excitation-inhibition balance is established only for some mitral cells leading to sparse responses of the mitral cell. Our model suggests a functional significance to the dendrodendritic synapses that mediate interactions between mitral and granule cells. The model accounts for the sparse olfactory code in the steady state and predicts that transient dynamics may be less sparse. INTRODUCTION Mammalian olfactory bulb the in f irst stage the is processing of information about odorants. The surface of olfactory bulb is covered by several thousands of glomeruli. Each glomerulus receives inputs from a set of sensory neurons expressing the same type of olfactory receptor protein. The inputs into each glomerulus are therefore substantially correlated (Shepherd et al., 2004; Lledo et al., 2005). The glomerulus-based modularity is preserved further by the mitral cells. Most of them receive direct excitatory inputs from a single glomerulus only. Mitral cells are also a major output class of the olfactory bulb that carries information about odorants to olfactory cortex (Figure 1). In 1950, Adrian reported that the spontaneous activ ity of the mitral cells in awake rabbit is much higher than that in an anesthetized animal, and in awake animal odor responses vanishe on the background of this spontaneous activ ity (Adrian, 1950). These observations were confirmed and more thoroughly analyzed in mice in recent study by Rinberg et.al. (Rinberg et al., 2006). The odor responses of mitral cells are sparse and state-dependent (Kay and Laurent, 1999; Rinberg et al., 2006; Fuentes et al., 2008). Only a small fraction of mitral cells responds to the indiv idual odorants despite receiv ing substantial odorant related input. How can sparseness emerge mechanistically? One of the mechanisms for the generation of sparse responses involves cells' f iring near threshold (Rozell et al., 2008). If the inputs into a cell fall below threshold, the cell does not fire, which leads to the sparse responses. However, this explanation does not apply in the case of mitral cells, because they exhibit high spontaneous firing rates in the awake behav ing animals, i.e. specifically in the state when the responses are most sparse. It is common to explain sparse codes as satisfying the constraint of metabolic energy efficiency (Lennie, 2003; Olshausen and Field, 2004). Although the information content of sparse representations is poor compared to the full codes (Baraniuk, 2007), sparse codes can use less metabolic energy. However, the high spontaneous rates coexisting with sparse representations makes the olfactory code both informationally poor and metabolically costly. The functional significance of sparse olfactory codes is therefore unclear. These arguments point to neurons other than m itral cells as likely sources of sparseness. Amongst the bulbar neurons, the granule cells (GC) are thought to implement lateral inhibition between m itral cells belonging to different glomeruli through a mechanism based on dendro-dendritic reciprocal synapses (Figure 1) (Shepherd et al., 2004). Such interactions are thought to facilitate discriminations between similar stimuli and mediate competitive interactions between coactive neurons (Arev ian et al., 2008). Because GCs are the most abundant cell type of the olfactory bulb it is likely that they are substant ially involved in these tasks. In this study we propose a novel role for the GCs of the olfactory bulb. W e show that granule cells can form complimentary to mitral cells representations of olfactory stimuli. Exact balance between the excitation from receptor neurons and the inhibition from the granule cells eliminates odor responses in some mitral cells. However, some mitral cells retain the ability to respond to odors, due 1 to incompleteness of the granule cells representation. The olfactory code on the level of mitral cells becomes sparse. This function is facilitated by the network architecture based on dendrodendritic reciprocal synapses between the mitral and granule cells. A Glomeruli Mitral cells Granule cells miW imW ig mx my B Granule to mitral cell synaptic weights: W G ranule cell index x e d n i l l e c l a r t i M Figure 1. Mitral – granule cell network model. A. Mitral cells (gray triangles) receive excitatory inputs, mx , from glomeruli, (large circles). The mitral cell output, my , is sent to the brain. Mitral cells and granule cells (small solid circles) are connected by reciprocal dendrodendritic synapses. Granule cells receive the excitatory input from the mitral cells, which defined by synaptic imW , and m itral cells are inhibited by granule cells, weight matrix miW . B. Granule to mitral cell synaptic weight with weight matrix miW . Here we argue that GCs represent mitral cell inputs as a superposition of columns in the weight matrix, such as the one marked by the dotted rectangle. RESULTS Our goal is to describe the behav ior of two ensembles of neurons: mitral and granule cells (Figure 1A). Mitral cells are projection neurons that receive inputs from olfactory glomeruli and send information about odorants to olfactory cortices for further processing. Mitral cell inputs from the mx (here and below glomeruli are contained in vector 2 index m represents the m itral cell number running between 1 and M , the total number of mitral cells). These inputs represent synaptic inputs from the receptor neurons. The output firing rate of the mitral cells is denoted by my . The purpose of the present study is to understand the relationship between m itral cell inputs mx and their outputs my in the presence of granule cell inhibition. interneurons. GC responses are GCs are inhibitory ia with the index i running described by the vector between 1 and N , the number of granule cells. We assume that the granule cells are much more abundant than mitral cells, i.e. N M (Tolley and Bedi, 1994). The mitral cells receive inhibitory inputs from the granule cells miW (Figure 1B). Each determined by the weight matrix miW describes the strength of synapses element of matrix between a granule cell number i and a mitral cell number m . Because of the relative abundance of granule cells, this matrix is rectangular as shown schematically in Figure 1B. Similarly, the reverse synaptic weights from the mitral imW to granule cells are described by the matrix . Two imW miW and reciprocal weight matrices, , although attributed to the same dendrodendritic synapse, are not necessarily the same. The system of mitral cells and granule cells connected by reciprocal dendro-dendritic synapses can be described by a Lyapunov function. The behav ior of the network can be simplified if the synaptic weights from imW mitral cells to granule cells are proportional to the miW : reverse synaptic weights from granule cells This assumption is not unrealistic because both weights are determined by active zones within the same dendrodendritic synapse (Shepherd et al., 2007), and can, in principle, be regulated to be proportional. In the Methods section we show that under this condition the dynamics of the network is described by minimization of the cost function that is called Lyapunov function (Hertz et al., 1991). Lyapunov function is a standard construct in the neural network theory that has been extensively used to study the properties of complex networks. The steady states of the neural activ ities can be obtained as m inima of this function. Thus, by f inding minima of the Lyapunov function one can understand the steady-state responses of W  im (1) W mi .     . mitral and GCs to odorants. The Lyapunov function for mitral - GC network has the following form: 2 1 M N        2     i i m 1 1  W a mi i  C a i (2) x m  The first term in the function contains the sum of squared differences between the excitatory inputs to the mitral cells from receptor neurons mx and the inhibitory inputs from the granule cells. The inhibitory inputs are proportional to ia weighted by synaptic the activ ity of granules cells miW . The f irst term therefore reflects the balance matrix between excitation and inhibition on the inputs to mitral cells. Minimization of this term due to the system 's dynamics leads to the establishment of an exact balance between excitation and inhibition. The second term of the Lyapunov function represents a cost imposed on the ia  , which 0 granule cell firing. It reaches m inima when correspond to minimization of number of active GCs. Granule cells represent mitral cell inputs as a superposition of dendrodendritic weights. W e will first assume that the second term in the Lyapunov function is negligible. This condition corresponds to the case when the granule cells are easy to activate and, therefore, the cost associated with their firing is small. Minimizing the first term of the Lyapunov function without the cost leads to the mx precise balance between excitation and W a [see equation (2)]: inhibition i mi i   i According to this equation the dendrodendritic weights can  iW be v iewed as vectors defined in the space of mitral cell inputs (gray arrows in Figure 2A). The firing rates of ia can be v iewed as unknown coefficients. granule cells The problem posed by equation (3) can be restated as follows: Represent the vector of mitral cell excitatory inputs  x iW as a superposition of vectors by f inding the ia (Figure 2). We conclude that the network coefficients dynamics in olfactory bulb leads to the representation of the bulbar inputs by the granule cells using the dendrodendritic synaptic weights as basis vectors. In this model the granule cells play the role of coding cells that carry representations of mitral cell inputs. W a mi i x m (3) . 2x x  iW 1x Figure 2. The olfactory system with two mitral and five GCs. The inputs into mitral cells from the receptor neurons are represented by the 2D vector x . The weights from the GCs are shown by five  iW 2D vectors . Equation (3) leads to the representation of the input vector in terms of the weight vectors, as shown by the dashed lines, with the coefficients given by the activities of the GCs. In this example, the second and the third GCs are active, although other solutions are also possible.  iW is equal to the The number of the basis vectors number of granule cells, N . The dimensionality of the input vector x is equal to a number of the mitral cells, M , which is much smaller than N . Therefore, the problem of representation of mitral cell inputs as a superposition of basis vectors may have several solutions. Thus, in Figure 2 f ive granule cells (gray arrows) have to represent inputs 1x and 2x ). Clearly there are many into two mitral cells ( ways how a 2D vector can be decomposed into five components. This picture suggests the following interpretation of the role of the bulbar GC. The strength of dendrodendritic synapses represents the receptive fields of indiv idual GCs, i.e. specific features in the glomerular inputs that these cells respond to. The bulbar network attempts to decompose its inputs into a superposition of these indiv idual features. Because of the abundance of the GCs, no unique representation of the inputs in terms of GC receptive fields can be found. The olfactory code as defined by the GCs is therefore underdetermined. As suggested underdeterm ined of studies the by (overcomplete) systems the unique solution can be found if one includes the cost imposed on the f iring rate vector into consideration. This is the second term in equation (2) that we so far neglected. Mitral cell firing rates result from the balance between excitation and inhibition. In our model the response of a mitral cell is defined as the dev iation of the firing rate of a ym : ym , rate, spontaneous from cell, mitral 3 ym  ym  ym . This response results from the excitatory inputs from receptor neurons mx and inhibitory inputs from the granule cells . (4) y   m N   x W a m mi i i 1  my of the mitral cell is a result of Therefore the response the balance between excitatory and inhibitory inputs. If such a balance is established precisely, i.e. if the granule cells represent exactly the bulbar inputs according to equation (3), no mitral cell is expected to respond to odorants. On one hand, this observation implies that the responses of mitral cells are ultimately sparse - they completely disregard their excitatory inputs. This model therefore contains the feature of the olfactory code observed experimentally, i.e., sparseness. Sparseness in this model emerges as a result of self-organizing balance between excitation and inhibition. On the other hand, it is hard to imagine a utility of this type of code, because the mitral cells do not respond to odorants at all. Next we identify the features of the model that could allow mitral cells to respond to odorants. Sparse mitral cell responses result from incomplete granule cell representations. How can responses of mitral cells emerge in our model? If GCs represent the excitatory inputs of the mitral cells exactly, mitral cells are unresponsive. Therefore GCs have to fail to represent glomerular inputs exactly for the mitral cells to respond to odorants. The GC code therefore has to be incomplete leading to an inaccurate odorant representation by the GCs. Here we present the simplest example of such an incomplete GC code. In Figure 3 we illustrate the synaptic currents emerging in the simplistic olfactory system with six mitral cells and only one GC. Vector x represents the excitatory inputs received by all six mitral cells. In Figure 3A the odorant supplies inputs into three m itral cells out of six. We then assume that the granule-mitral cell weights W match exactly this input pattern. As a result, when the   into to GC is activated, it will return an inhibitory input x the mitral cells that exactly matches the pattern of excitation. The responses on the m itral cells to the odorant will be very weak as a result of such a compensation (some small residual responses are needed to drive the GC). Figures 3B and C show examples of two odorants that activate similar subsets of mitral cells. The pattern of   however is exactly the same as in Figure 3A, inhibition x because with the single GC present, the inhibitory currents are constrained to a single pattern. As a result of subtraction between receptor neuron excitatory and GC inhibitory inputs into the mitral cells, only a single mitral cell responds to each of these odorants. Thus, the responses of mitral cells become sparse. Some mitral cells relay their inputs, while others are almost completely inhibited by the GCs. That some mitral cells respond is a consequence of the incompleteness of the GC code. Mitral cell input x Granule cell synapses W Granule cel l to mitral cel l inhibition   x Mitral ce ll response excit.-inhib.     x x y    A r e b m u n l l e c l a r t i M B C D Figure 3. Simple olfactory system that contains six mitral and one granule cell. Gray areas indicate strong activity. Light gray areas on the right indicate weak responses due to possible small correction induced by the cost function. (A-D) responses to four different odorants represented by the combinatorial inputs shown on the left. The result of sparsening of the m itral cell responses is the reduction of redundancy of the odorant representation. Thus, although m itral cells may receive highly redundant inputs (Figure 3B and C, lef t column), only few may respond (right column). The sparseness in the mitral cell responses may be important for reducing overlaps between odorant representations. in the example In Figures 3B and C, responses of mitral cells to two odorants share no overlap. Figure 3D shows the responses to an odorant that are inhibitory (below the baseline activ ity). Therefore, the incompleteness in the GC odorant representation may emerge from the restriction on the diversity in the GC weights. 4 Granule cell nonlinear input-output relationship may be responsible for sparse granule cell coding. The example presented above contained only one GC. Therefore GC code was incomplete leading to the mitral cell firing. If the number of GCs was equal or larger than the number of mitral cells, there are enough vectors in the basis to represent any glomerular input. Because the olfactory bulb contains many more GC than mitral cells one should seek some other source for the incompleteness of GC representation. We propose that a source of incompleteness is in the constraint that GC firing rates cannot be negative. To satisfy non-negativ ity condition for the GCs firing rate, iC a in equation (2) should be infinite  the cost function  iC a is defined by the for negative ai . In general, cost non-linearity in the cells’ input-output relationship. The simplest example is shown at Figure 4. For the positive firing rates the cost function increases proportionally to the GC threshold for firing , making it more costly to fire for the cells with a large threshold. This feature selects which GC will be active in the stationary state with the higher threshold cells less likely to fire. The exact relationship between the input-output relationship and the cost function is given by equation (11) of the Methods. A a ( )g u B ( )C a a a 0  u Figure 4. Granule cell input-output relationships (A) and the ( )C a (B). u is the granule cell's corresponding cost-functions input current. a is the resulting firing rate. The cost function prohibits negative firing rates, which reflects the feature of the input-output relationship. For positive firing rates the cost-function is approximately linear. Because GC firing rates cannot be negative, some odorants cannot be exactly represented by the GCs. Indeed, consider the gray regions in Figure 2A. The input vector x in this area cannot be represented as a superposition of GC weights (gray arrows) with posi tive coefficients. To exactly represent a vector in this area, at least one GC has to have a negative firing rate, which is not allowed. Therefore, the inputs within gray area will lead to an incomplete representation by the granule cells. The exact balance between inhibition and excitation on the input of some mitral cells will be destroyed and, according to equation (4), some mitral cells will respond to these inputs. A more high-dimensional example is shown in Figure 5. From this example it is clear that substantial responses of mitral cells emerge when the input from receptor neurons falls outside the conical domain spanned by the GC weight vectors (gray vectors in Figure 5B). In this case the inhibitory currents from the granule cells to the m itral cells   on conical domain that are given by the nearest vector x can be represented as a superposition of GC weights with positive coefficients. The response of the mitral cells to the odorant it the result of the difference between excitatory receptor neuron inputs x and inhibitory inputs from the   . Thus, substantial mitral cell response is granule cells x    , i.e. there is no precise balance of possible if x x inhibition and excitation on the input to mitral cells. In this case the representation of the receptor neuron inputs by the GC is incomplete. The responses of GC are sparse for the incomplete representations. Thus, in Figure 5B, for the 3D space characterizing inputs into three mitral cells only two GCs out of eight (red arrows) are active. This is because GCs approx imate the vector lying on the surface of the region spanned by GC weights that is 2D. In multidimensional space, when M mitral cells are present, the active GCs weight vectors lie on the surface of the domain spanned by all GCs, so the number of active GCs is smaller or equal 1M  . Because the number of GCs is much larger that the number of mitral cells N M that , one expects responses of GCs are very sparse. This reasoning is confirmed more rigorously in the Theorem 1 of Methods. Complete GC representation A  x   x Incomplete GC representation  B   x y x  iW  iW Figure 5. The olfactory bulb with three mitral cells (3D input space) and eight GCs. The inputs into mitral cells from receptor neurons are represented by the black arrows. The synaptic weights from eight GCs to the mitral cells are shown by the gray arrows. (A) If inputs lie within the cone restricted by the weight vectors (dashed) the input vector can be represented as a superposition of weight vectors exactly. The GC representation of 5 To demonstrate our point we will use simplified model of bulbar network containing only one granule cell (Figure 3) This network has an advantage that an exact solution can be found and understood even when a f inite threshold for firing is present in the activation of the GCs. Indeed, consider input configuration shown in Figure 3B. Assume for simplicity that all of the non-zero weights and mitral cell inputs have unit strengths. Then, the Lyapunov function for the activ ity of the single GC a is K a a ( ) (1    . (5) 2 a  . K is the 0 Here, of course, we have to assume that 3K  in Figure number of non-zero weights for the GC ( 3). By minimizing the Lyapunov function we obtain a K  K  1 / , for , and zero otherwise. The activ ity of the mitral cell number two (Figure 3B and 6) cannot be affected by the GC, because the latter makes no synapses onto this cell. The activ ities of mitral cells 1, 4, and 6 are given by a  ) 2 (6) y  1, 4,6 a 1     K K  for . This formula means that the mitral cells will increase their firing rate slightly to activate the GC. The amount of increase is equal to the threshold for activation of the GC div ided by the number of mitral cells contributing to the input current, i.e. K . The activ ity of the GC is also assumed to raise fast above the threshold so that it suppresses all firing increases on its input above this value [equation (6)]. The responses of m itral cells as functions of the threshold  are shown in Figure 6. Figure 6 shows that for large thresholds  all mitral cells that receive receptor inputs will respond to the odorant. In this regime the GC is silent due to the high threshold and, therefore, the mitral cells relay the receptor neuron inputs directly. The m itral cell code is not sparse in this case. W e argue that the case of large thresholds corresponds to the network in the anesthetized animal. When the threshold for GC firing is small, the m itral cell firing becomes sparse (Figure 6). This is the regime considered throughout this paper, which, we argue, corresponds to the awake animal. According to this model therefore the transition from awake to anesthetized state is accomplished by an increase in the thresholds of GC firing, which could be mediated by the centrifugal projections. mitral cell inputs is exact and the mitral cells are expected to respond weakly due to exact balance between inhibition and excitation. (B) If the input vector is outside the cone of GC weights (dashed) the GCs cannot represent the inputs precisely. This is because the GC firing rates (coefficients of expansion) cannot be negative. The best approximation of the inputs is    is the nearest vector on the cone  . Vector x shown by vector x to the input vector x . This vector is formed by two GCs (red).  Because x  , the GC code is differs from GC representation x representation Incompleteness of incomp lete. to leads     . substantial mitral cell firing y x x    Incomplete granule cell representations are typical for random network weights. It may appear that incomplete representations are lim ited to the edges of the parameter space, such as the gray area in Figure 2A. The situation shown in Figure 5A may seem to be more typical than that in Figure 5B. Here we argue that this intuition applies only to input spaces of small dimensionality such as shown in Figures 2 and 5. For a multidimensional input space, which corresponds to large number of m itral cells, random inputs and network weights create incomplete representations. This implies that it becomes almost impossible to expand a random input vector with positive components into the basis containing vectors with positive components using only non-negative coefficients without the loss of precision. In the Methods section we show that the number of coactive GC for random binary inputs with M mitral cells is ~ M . Because a precise representation of the M - dimensional random input requires M vectors, this result implies that the representation of odorants by the GCs is typically imprecise. The GC code is therefore incomplete. for sparse GC-to-mitral cell We also show that connectiv ity, when only K M weights are non-zero, the number of coactive GC larger is somewhat /M K . This number is still substantially smaller that ~ what is required for complete GC code. We conclude that GC cannot represent m itral cell inputs precisely for the case of random connectiv ity, which implies ubiquity of incomplete representations. The state dependence of the granule cell code. Above we have discussed two reasons for the incompleteness of GC representation. First, the incompleteness may result from the restricted diversity of the GC weights (Figure 3). Second, we suggested that the non-negativ ity of the GC firing rates may lead to an inaccurate representation of odorants implying incompleteness. Here we discuss the third reason why GC may yield an inaccurate code. W e suggest that the presence of f inite (non-zero) threshold for GC firing (Figure 4A) can lead to an inaccurate representation. This phenomenon allows us to explain the transition between the awake and anesthetized responses. 6 1L norm is called sparse and respond to a stimulus. Thus, the corresponding neural representation is called sparse overcomplete. It was found that a network of inhibitory neurons can implement sparse overcomplete codes (Rozell et al., 2008). To show this, the dynam ics of the system was represented as m inimization of a cost function called a Lyapunov function. The basic idea of this method goes back to John Hopf ield, who was the f irst to evaluate the Lyapunov function for a recurrent network (Hopf ield, 1982, 1984). Hopfield network has attractor states that allow to implement associative memory (Hopf ield, 1982; Hertz et al., 1991). In contrast to Hopfield networks, in purely inhibitory networks, the recurrent weights entered the Lyapunov function with a m inus sign, which abolished the attractor memory states and made the network purely sensory (Rozell et al., 2008). Minimization of Lyapunov function in realistic recurrent networks with inhibition was suggested to implement the parsimony constraint mentioned above. To implement sparse overcomplete representations with realistic networks of neurons at least two requirements on the network connectiv ity have to be met (Rozell et al., 2008). First, the feedforward weights between the input layer of the network and the inhibitory neurons have to contain the dictionary elements (Figure 7A). Thus, the inhibitory neurons that are responsible for a particular dictionary element will be driven strongly when it is present in the input, due to high overlap between the stimulus and the feedforward weight. Second, the recurrent inhibitory weight between any pair of neurons has to be proportional to the overlap between their dictionary elements (Figure 7A). Because inhibition between neurons implements competition, tuned that similarly implies feature this neurons compete most strongly. Therefore, the two types of network weights, feedforward and recurrent, have to be in close match, one of them constructed as the overlap of the other. It is not clear how this match is established in a biological system, because these are different sets of synapses. Here we argue that the olfactory bulb network architecture that is based on dendrodendritic synapses can ensure that the feedforward and recurrent connect iv ity are in close match. Mi tral cell responses y 2y 1, 4 , 6y 0 K  Figure 6. Sparsification of responses as a function of parameter  (threshold for GC firing). Example shown in Figure 3B is used. We argue that for large thresholds the response of mitral cells is less sparse than for small thresholds. Because the threshold for GC firing can be regulated by the centrifugal inputs from cortex, increase in firing threshold may mediate the transition between awake and anesthetized states. DISCUSSION Sparse overcomplete codes. At the core of these representations is an assumption that a sensory input can be decomposed into a superposition of primitives that are called dictionary elements (Field, 1994; Olshausen and Field, 2004; Rozell et al., 2008). The decomposition is sought in the form of a set of coefficients with which different dictionary elements contribute to the input. These coefficients are thought of as responses of neurons in a high level sensory area that indicate whether given feature is present in the stimulus or not. Because the number of primitives several large, usually is available decompositions are consistent with the inputs. If the set of primitives is overcomplete, the decomposition problem generally has no unique solution. To make sensory representation unambiguous some form of parsimony principle is added to the model in the form of cost function on the coefficients/responses. The solution that yields a minimum of the cost function is assumed to be chosen by the nervous system. The decomposition is found to be dependent on the cost function. The general form of a cost in power : is a sum of function firing rates  . For L2 , a simple sum of squared  L  ai i coefficients, it can be shown that generally all neurons will respond to any stimulus, and, therefoore, the code is not 1L norm when the cost is equal to the sum of sparse. For absolute values of the responses, the solution is found to be sparse. This means that only a few neurons will 7 Sparse overcomplete representation networks Recurrent layer A Input layer imd G ik ?   m d d im km Olfactory bulb B Mitral cells Granule cells imd G ik   m d d im km Figure 7. A possible functional significance of dendro-dendritic synapses. (A) Networks implementing sparse overcomplete representations. These networks decompose the input pattern into a parsimonious superposition of dictionary elements. Response of each cell in the recurrent layer indicates whether given dictionary element is present in the input. The feedforward excitatory weights contain the dictionary elements. The recurrent inhibitory weights are proportional to the overlap between dictionary elements. More similar input patterns compete stronger. It is not clear however how this condition on the recurrent weights is implemented biologically. (B) In the networks with dendrodendritic synapses the connectivity between granule cells is inhibitory. Because every two granule cells are two synapses away from each other, the strength of pair wise inhibition contains overlap between dendrodendritic weights thus automatically satisfying the constraint on the recurrent inhibitory weights shown in (A). Dendrodendritic synapses facilitate the representation of odorants by the granule cells. GCs of the olfactory bulb receive excitatory inputs from the mitral cells through the dendrodendritic synapses (Shepherd et al., 2004). These synapses encode patterns that can drive strongly indiv idual GCs. At the same time the effective connectiv ity between GCs is inhibitory (granule to mitral and m itral to granule synapses are inhibitory and excitatory respectively). Here we argue that the strength of mutual inhibition between GCs may be proportional to the overlap between their input weights. The reason for this is that indiv idual GCs are two synapses away from each other. To calculate their mutual inhibition one has to sum over all intermediate synapses, which leads to evaluation of a convolution or overlap (Figure 7B). Therefore, GCs inhibit each other stronger if their inputs from mitral cells are 8 to more sim ilar. Thus, both requirements necessary implement sparse overcomplete representations stated prev iously are met. This implies that the function of GCs may be to detect specific patterns of activ ity in the inputs that m itral cells receive. The GCs then build the parsimonious representation of mitral cell inputs. The parsimony is ensured by the mutual inhibition between GCs, with more sim ilar GCs inhibiting each other stronger. The latter condition is facilitated by the network architecture based on dendrodendritic synapses. This observation prov ides a potential explanation the for existence of these synapses that so far have been discovered in the olfactory bulb only (Shepherd et al., 2004). Mitral cells respond to odorants because granule cell code is incomplete. Two problems emerge if we assume that granule cells implement sparse overcomplete codes. First, granule cells are interneurons and, as such, cannot directly transmit their representation to the upstream network. The significance of these representations becomes unclear. Second, we find that if granule cells establish an absolutely accurate representation of their inputs, the mitral cells will respond to odorants very weakly. This is because granule cells can eliminate these responses from the mitral cells' firing. These considerations suggested to us that the representations by the granule cells are incomplete. This implies that granule cells cannot find accurate representations of their inputs, for example, because this would require for the firing rates to become negative. If granule cells' codes are incomplete the m itral cells transmit only the unfinished portion of the representation to the upstream olfactory networks. As a consequence the representations of odorant by the m itral cells becomes sparse. The redundancies in the m itral cell codes are reduced and the overlaps in representations of similar odorants are erased. As a result the patterns of mitral responses to similar odorants become more distinguishable from each other. The Lyapunov function. Lyapunov functions are standard tools in the neural network theory (Hertz et al., 1991). (Seung et al., 1998) have shown that the network that contains two populations of neurons, inhibitory and excitatory, can be described by the Lyapunov function. This model can be mapped onto the system of mitral-GCs. Here we have eliminated the dynamics of mitral cells from consideration and reduced the description to GCs only. By doing so we have shown that GC could be used to extract features from mitral cell inputs. (Lee and Seung, 1997) have proposed a network mechanism that implements conic encoders, i.e. the ones that represent inputs under non-negativ ity constraint. Although the non-linear network mechanism is somewhat different from the one used here, the set of encoding/error neurons in Lee and Seung can be v iewed as analogs of granule/mitral cells, respectively, in our model. In our study we propose a mapping of the conic networks onto the olfactory bulb network and study the implications of this mapping for the olfactory code. Feedback from higher brain areas and state dependence of the olfactory code. The sparseness of the mitral cell responses depends on the nonlinearity of the GCs, and specifically on the GC activation threshold . In this study we assumed that GCs have similar activation threshold that is small enough for them to be easily activated by small levels of activ ity in the mitral cells. If the thresholds for activation of indiv idual GCs are different, one can env ision a mechanism how olfactory code carried by both m itral and granule cells can be controlled to adapt to a particular task. Thus, if the threshold for activation is raised for a subset of GCs, these cells will no longer be active and, therefore, their activ ity will not be extracted from the firing of mitral cells. If, for example, the threshold for all of the GCs is increased thus making them unresponsive, the olfactory code carried by the mitral cells replicates their inputs from receptor neurons. Conversely, if the activation threshold is lowered for a subset of GCs, these cells will efficiently extract their activ ity from the m itral cells' responses. Thus, a particular redundancy amongst similar odorants can be excluded in a task-dependent manner. Therefore, the thresholds for GC activation may regulate both an overall sparseness of mitral cell responses and the f ine structure of the bulbar olfactory code. GC threshold values depend on cellular properties, but also can be effectively modulated by additional input into these cells. GCs in the mammalian olfactory bulb are recipients of the signals from the efferent projections from the cortex and other brain areas (Dav is and Macrides, 1981; Luskin and Price, 1983). Additional signal to GCs can change the effective threshold values. If a GC receives excitatory inputs from the cortex, the m itral cell signal is closer to the threshold value and the granule cell is easier to excite by the odorant-related inputs. These features prov ide an opportunity to control sparseness and information content of the mitral cell signal by regulating the input to the GCs from the cortex. Implicit experimental ev idence for the presence of such regulation already exists. In the anesthetized state, when the efferent signal coming to the bulb from the cortex is minimal, the mitral cells respond strongly to the odor stimulation. In the awake state, when the cortex is active, the mitral cell code becomes much sparser (Adrian, 1950; Rinberg et al., 2006). Cutting lateral olfactory tract and eliminating feedback from the brain in awake rabbit made mitral cells respond in the similar manner as in an anesthetized rabbit (Moulton, 1963). Therefore centrifugal projection may indeed regulate sparseness of the olfactory code. Some ev idence also points towards the possibility of finer network tuning by specific activation or deactivation of to enhance extraction of relevant subsets of GCs information. First, (Doucette and Restrepo, 2008) demonstrated that the mitral cell responses to odorants change as animals learn the task. Second, Fuentes et.al. show that the number of mitral cell responses depends on the task (Fuentes et al., 2008). When a rat is involved in odor discrimination task the average number of mitral cell excitatory responses is less then in the rat passively smelling an odor. The assumption is that when an animal discriminates odorants, to it may be advantageous suppress redundant m itral cell responses and enhance those, which carry the most behav iorally relevant information. Our model proposes the neuronal mechanism for this phenomenon. Transient regime. In this study we explicitly considered the steady state regime and used Lyapunov function approach to find the response of the mitral and GCs to the stimulus in the stationary state. Obv iously the system 's response to odorants is dynamic. What happens when the stimulus appears, when an animal first sniffs an odorant? The thorough analysis of this problem is beyond the scope of this paper, however, we would like to present some obv ious points, which lead to specif ic experimental predictions. If the signal first appears on the input to mitral cells, it causes elevation of the mitral cell firing rates, which in turn causes activation of GCs and suppression of m itral cells by feedback inhibition. The initial responses of the mitral cells are therefore represented by quick bursts of activ ity followed by decay to the steady state described in this study. The olfactory bulb may transmit information about the stimulus by these quick bursts of activ ity. This observation leads to several questions. First, what information about the stimulus is sent to the cortex by transient bursts and of activ ity and the consequent steady state? Second, what are the experimental conditions for observations of such bursts? While it is hard to say conv incingly anything about the role of different modes of information transmission, we may speculate about the second question. The mitral cell bursts of activ ity presumably are short and stand on top of the high spontaneous activ ity. In order to observe them to synchronize spike trains with reliably one needs stimulus delivery. In mammals, stimulus delivery is controlled by sniffing. In the prev iously studies (Kay and Laurent, 1999; Rinberg et al., 2006; Doucette and Restrepo, 2008; Fuentes et al., 2008), the authors 9 synchronized their recordings with stimulus onset, but not with sniffing/breathing. This approach may to lead smearing of the short bursts, and emphasize the steady state responses of the network. In such regime the odor responses should be sparse as predicted by the model. New ev idence by (Shusterman et al., 2009) and (Cury and Uchida, 2009) suggests the presence of fast bursts of mitral cell activ ity synchronized with sniffing. Conclusions. In this study we addressed theoretically the experimental ev idence of sparse odor codes carried by the mitral cells in awake behav ing rodents. W e proposed a novel role for the granule cells of the olfactory bulb in which they collectively build an incomplete representation of the odorants in the inhibitory currents that they return to the m itral cells. Because the representation formed by granule cells is incomplete, mitral cells can carry information to the cortex in the form of sparse codes. This function is facilitated by the network architecture that includes bidirectional dendrodendritic mitral-to-granule cell synapses. W e suggest that the synaptic strengths in the indiv idual synapses are correlated between the two directions. Acknowledgments. The authors are grateful to Dmitry Chklovskii, Barak Pearlmutter, and Sebastian Seung for helpful discussions. METHODS Derivation of the Lyapunov function. Our model is based on the following equations describing the responses of mitral and granule cells respectively: N   y x W a   m m mi i i 1  M   W y im m 1  g u a ( )  (9) , i i iu is the membrane voltage of the granule cell and where y y y    . We use the following abbrev iation for the m m m u  . We assume therefore /du dt temporal derivatives that the dynamics of mitral cell responses is fast enough to reflect the instantaneous values of inputs. If one def ines a function of granule cell firing rates  1 M    a ( ) (cid:0)  2    i m 1 where the cost function is defined as a     a da g C a ') 0 W a mi i  C a i N  i 1  u   , (10) (11) (7) (8) x m     u     ,   1 ( m m m 2 ' 2    0 (12) one can show that the behav ior or the system can be v iewed as i.e. form of gradient descent, a ui   / ai . The time derivative of the Lyapunov function  is therefore always negative     ai   ( ui ) ai ai i i      ai g1 ( ai )        i ( )g u is a monotonically The last inequality holds if increasing function. Therefore the Lyapunov function always decreases if the system behaves according to equations (7) through (9). Since the function is limited from below, the stable states of the system are described by the minima of the Lyapunov function. threshold-linear neuronal the function for The cost u   g u ] [ ( )   activation function shown in Figure 4A can be approximated by a linear function when the firing rate of the granule cell are not too large a a C a 0 ( ) ,  (13) For negative values of firing rates a the cost function is infinitely big, reflecting the fact that negative firing rates are not available. The importance of condition (1) For our model to have a Lyapunov function a more general condition than (1) may hold. Indeed, the sufficient condition for the Lyapunov function to ex ist is that the WimWmk  Gik  network weight matrix is symmetric  i (Hertz et al., 1991). This will be true if , for example, W E  nmE is an arbitrary symmetric M W where mi nm in by M matrix. Thus condition (1) is sufficient but not necessary for the network to have a Lyapunov function. We argue however that this condition is necessary for the system to be described by the Lyapunov function in the form of equation (2). The number of coactive granule cells Here we will distinguish two types of Lyapunov functions. First, we will consider the homogeneous case, when the Lyapunov function has the form  1 M     a ( ) (cid:0)   0 2     i m 1 In the homogeneous case we disregard by the contribution of the cost function other than constraining the firing rates to be non-negative. This condition corresponds to vanishingly small threshold for the activation of the granule cells. The non-homogeneous Lyapunov function is W a mi i (14) x m   . 2 10 (15)     a a a ( ) ( ) (cid:0) (cid:0)  i i 0 i ia  . We will prove here two 0 with the same constraint theorems that limit from above the number of coactive ia  ). A quantity 0 granule cells (i.e. the ones for which that will be important to us is the error in approx imation   W a x x   that is also, according to equation i mi m m i (4), the response of mitral cells. Theorem 1: Sparseness of the homogeneous solution Assume that the M -dimensional receptive fields of the  iW are the vectors of general position, i.e. granule cells  every subset of M vectors iW are linearly independent. Then, in the m inimum of the homogeneous Lyapunov mx  for all m (i.e. mitral cells do 0 function (14), either not respond, granule cell representation is complete) or fewer that M granule cells are active. In the former case mx  ) all of the granule cells may be active. 0 ( Proof: assume that M or more GCs are simultaneously active. Let us vary slightly the activ ity of only one GC: ka    . The corresponding variation in the Lyapunov     x W O ( ) 2 (cid:0)       . Because we are function is k 0 considering the minimum of the Lyapunov function, all of   the scalar products  k x W  have to be zero, which is   kW x  0 or the number of vectors possible only if less than M . Theorem 2: Sparseness of the inhomogeneous solution the set of N M -dimensional vectors Consider     iW  ,   . Assume that these vectors are of general i i 1M  of these vectors is position, i.e. any subset of linearly independent. Then in the minimum of the inhomogeneous Lyapunov function (15) no more than M granule cells can be simultaneously active. Proof: For any active granule cell the following equation is valid:  a W x  / (cid:0)      i mi m i m Assume that more than M granule cells are active. Then 1M  such equations for M unknowns we have at least mx . Such a system 1M  in general case (if (16) is   0 . 11  i corresponding vectors are independent) is inconsistent and has no solution. Thus the number of coactive granule cells cannot exceed M . Note: Consider the case of small but non-zero thresholds of firing of granule cells . In this case two regimes can be distinguished. If vector x can be expanded in terms of  with positive coefficients, the f iring rates of M iW vectors granule cells are generally ~ 1 but the responses of m itral cells are small ( ~ ). This is the regime of sparse overcomplete representation. If the glomerular input vector x cannot be represented as a superposition of granule  iW cell weights with positive coefficients (incomplete representation), the responses of cells are essent ially (ignoring contributions ~ ) given by the solution of homogeneous problem (14), which explains our attention to this problem. In this case, according to Theorem 1, fewer than M GCs have large firing rates and only one has a small firing rate ( ~ ). Coactive granule cells the case of sparse in connectivity. Assume that the weight matrix between granule and mitral cells is sparse, which implies that out of miW for the i -th GC only K M M matrix elements are non-zero. This corresponds to the situation when the GC can contact a limited subset of only K m itral cells. Here we will calculate the number of GCs that are coactive in the case of sparse connectiv ity. Clearly, fewer than M GCs have to be coactive according to Theorem 2. W e argue however that the number of coactive GC can be substantially smaller than M . 0 i  (16) leads to the following approx imate W ith ia -s equation that is only valid for non-zero  Ka K M a p ( / ) 2  , j i j   p W x   where is the projection of the input vector on i i synaptic weight. W e can find therefore the activ ities of the GCs approx imately as functions of the threshold T :   . W e S a T SK M p T K a 2 / ( ) /    , , where i i i i can replace summation in the last equation with integration of distribution the over projections, K N p p K K ( exp[ ( ) ) / ] / 2       (we will assume that the sparseness of input vectors is 50% for simplicity). For T K K   we obtain (17) i Lennie P (2003) The cost of cortical computation. Curr Biol 13:493-497. Lledo PM, Gheusi G, Vincent JD (2005) Information processing in the mammalian olfactory system. Physiol Rev 85:281-317. Luskin MB, Price JL (1983) The topographic organization of associational fibers of the olfactory system in the rat, including centrifugal fibers to the olfactory bulb. J Comp Neurol 216:264-291. Moulton DG (1963) Electrical activ ity in the olfactory system of rabbits with indwelling electrodes. In: Olfaction and Taste I (Zotterman Y, ed), pp 71-84. Oxford: Pergamon Press. Olshausen BA, Field DJ (2004) Sparse coding of sensory inputs. Current Opinion in Neurobiology 14:481- 487. Rinberg D, Koulakov A, Gelperin A (2006) Sparse odor coding in awake behav ing mice. J Neurosci 26:8857-8865. Rozell CJ, Johnson DH, Baraniuk RG, Olshausen BA (2008) Sparse coding v ia thresholding and local competition in neural circuits. Neural Comput 20:2526-2563. Seung HS, Richardson TJ, Lagarias JC, Hopf ield JJ (1998) Minimax and Hamiltonian dynam ics of excitatory-inhibitory networks. In: Advances in neural information processing systems 10, pp 329 - 335. Shepherd GM, Chen WR, Greer CA (2004) Olfactory Bulb. In: The synaptic organization of the brain (Shepherd GM, ed), pp 165-216. New York: Oxford University Press. Shepherd GM, Chen WR, W illhite D, Migliore M, Greer CA (2007) The olfactory granule cell: from classical enigma to central role in olfactory processing. Brain Res Rev 55:373-382. Shusterman R, Koulakov A, Rinberg D (2009) The temporal structure of mitral/tufted cell odor responses in the olfactory bulb of the awake mouse. In: Society for Neuroscience annual meeting. Chicago. Tolley LK, Bedi KS (1994) Undernutrition during early life does not affect the number of granule cells in the rat olfactory bulb. J Comp Neurol 348:343-350.  (18) . (19) exp   K   ln N K / S T ( ) N K T K ) / ( 2   T K ( ) 2  This equation has to be solved self-consistently with T SK M 2 /  to obtain the threshold for the input projection T for granule cells to be active. Because the number of coactive granule cells can also be found from )p we have ( integrating   S T K M GN T ( ( ) )   REFERENCES Adrian ED (1950) The electrical activ ity of the mammalian olfactory bulb. EEG Clin Neurophysiol 2:377-388. Arev ian AC, Kapoor V, Urban NN (2008) Activ ity- dependent gating of lateral inhibition in the mouse olfactory bulb. Nat Neurosci 11:80-87. Baraniuk RG (2007) Compressive sensing. Ieee Signal Proc Mag 24:118-+. Cury K, Uchida (2009) Odor coding based on spike timing with respect to respiration-coupled oscillations during for In: Society sampling. active Neuroscience annual meeting. Chicago. Dav is BJ, Macrides F (1981) The organization of centrifugal projections from the anterior olfactory nucleus, ventral hippocampal rudiment, and piriform cortex to the main olfactory bulb in the hamster: an autoradiographic study. J Comp Neurol 203:475-493. Doucette W , Restrepo D (2008) Profound context- dependent plasticity of mitral cell responses in olfactory bulb. PLoS Biol 6:e258. Field D (1994) What is the goal of sensory processing? Neural Computation 6. Fuentes RA, Aguilar MI, Aylwin ML, Maldonado PE (2008) Neuronal activ ity of mitral-tuf ted cells in awake rats during passive and active odorant stimulation. J Neurophysiol 100:422-430. Hertz J, Krogh A, Palmer RG (1991) Introduction to the theory of neural computation. Cambridge, Mass.: Perseus. Hopfield JJ (1982) Neural networks and physical systems with emergent collective computational abilities. Proc Natl Acad Sci U S A 79:2554-2558. Hopfield JJ (1984) Neurons with graded response have collective computational properties like those of two-state neurons. Proc Natl Acad Sci U S A 81:3088-3092. Kay LM, Laurent G (1999) Odor- and context-dependent modulation of mitral cell activ ity in behav ing rats. Nat Neurosci 2:1003-1009. Lee DD, Seung HS (1997) Unsuperv ised learning by convex and conic coding. In: Advances in Neural Information Processing Systems 9 pp 515-521. 12
1311.3211
1
1311
2013-11-13T17:04:41
Stochastic inference with deterministic spiking neurons
[ "q-bio.NC", "cond-mat.dis-nn", "cs.NE", "physics.bio-ph", "stat.ML" ]
The seemingly stochastic transient dynamics of neocortical circuits observed in vivo have been hypothesized to represent a signature of ongoing stochastic inference. In vitro neurons, on the other hand, exhibit a highly deterministic response to various types of stimulation. We show that an ensemble of deterministic leaky integrate-and-fire neurons embedded in a spiking noisy environment can attain the correct firing statistics in order to sample from a well-defined target distribution. We provide an analytical derivation of the activation function on the single cell level; for recurrent networks, we examine convergence towards stationarity in computer simulations and demonstrate sample-based Bayesian inference in a mixed graphical model. This establishes a rigorous link between deterministic neuron models and functional stochastic dynamics on the network level.
q-bio.NC
q-bio
Stochastic inference with deterministic spiking neurons∗ Mihai A. Petrovici1*, Johannes Bill2*, Ilja Bytschok1, Johannes Schemmel1, Karlheinz Meier1 1Kirchhoff Institute for Physics, University of Heidelberg 2Institute for Theoretical Computer Science, Graz University of Technology (Dated: June 12, 2018) The seemingly stochastic transient dynamics of neocortical circuits observed in vivo have been hypothesized to represent a signature of ongoing stochastic inference. In vitro neurons, on the other hand, exhibit a highly deterministic response to various types of stimulation. We show that an ensemble of deterministic leaky integrate-and-fire neurons embedded in a spiking noisy environment can attain the correct firing statistics in order to sample from a well-defined target distribution. We provide an analytical derivation of the activation function on the single cell level; for recurrent networks, we examine convergence towards stationarity in computer simulations and demonstrate sample-based Bayesian inference in a mixed graphical model. This establishes a rigorous link between deterministic neuron models and functional stochastic dynamics on the network level. PACS numbers: xxx-xxx Introduction In responding to environmental sensory stimuli, brains have to deal with what is typically limited, noisy and ambiguous data. Based on such imperfect information, animals need to predict and react to changes in their environment. The recent hypothesis that the brain han- dles these challenges by performing Bayesian, rather than logical inference [1–3], has been strengthened by electro- physiological data which identified neural correlates of the involved computations [4, 5] and theoretical work on spiking network implementations [6–8]. In probabilistic inference, the potential values of a quantity are described by a random variable (RV) zk and all knowledge about dependencies between ran- dom variables is stored in a joint probability distri- bution p(z1, . . . , zK). The Bayesian belief about a set of unobserved RVs {z1, . . . , zM} given an observed set of RVs is represented by the posterior distribution p(z1, . . . , zM zM +1, . . . , zK). In particular, the posterior contains information not only on the most likely conclu- sion and potential alternatives, but also on the level of uncertainty of the outcome. Theoretical work [2] has argued in favor of sample- based representations of probability distributions in the brain. In this representation, instead of providing the entire distribution at any point in time, only samples z(t) ∼ p(z1, . . . , zK) are used as a proxy. When model- ing large systems, this offers three important advantages. First, approximate solutions can be provided at any time, with increasingly reliable results as the calculation pro- gresses ("anytime computing"). Secondly, in a sample- based representation marginalization comes at no cost, as p(zk) can be determined by simply neglecting the values of all other RVs. Thirdly, some sampling algorithms such as Gibbs sampling support a high degree of paralleliza- tion. In particular, the resulting algorithmic structure is reminiscent of neural networks [9]. FIG. 1. (A) Interpretation of spike patterns as samples of a binary random vector z. The variable zk is active for dura- tion τon (grey bar) after a spike of neuron k. (B) In stochastic neuron models, internal state variables modulate the instan- taneous firing probability (red). In contrast, deterministic integrate-and-fire neurons elicit a spike when the membrane potential crosses a threshold voltage (blue). The probability of firing as a function of the respective internal variable is represented by the grayscale in the background. Recently, a theory has been suggested which combines these advantages by implementing Markov chain Monte Carlo sampling in networks of abstract model neurons [7]. In this framework, spike patterns are interpreted as samples of binary RVs as follows (see Fig. 1A): k = 1 ⇔ Neuron k fired in (t − τon, t] z(t) (1) . The duration τon of the active state following a spike is a free parameter; in cortex, τon ≈ 10 ms is a good esti- mate for the timescale on which a spiking neuron affects the membrane potential of downstream cells. The neu- ron model underlying [7] is inherently stochastic (Fig. 1B top), with an instantaneous firing rate defined by rk(t) = lim ∆t→0 (cid:26) 1 p(spike in [t, t + ∆t)) = τ exp(vk) 0 ∆t if zk = 0 if zk = 1 , (2) where vk represents an abstract membrane potential. In contrast to this approach, in vitro experiments have demonstrated the largely deterministic nature of single neurons [10]. Similarly, microscopic models of neural cir- cuits typically rely on deterministic dynamics of their constituents. An often-used mechanistic model is the leaky integrate-and-fire (LIF) neuron rewrite (3) as τeff du dt = ueff − u , 2 (5) wki (Erev i − uk) δ(t− ts) , (4) τeff du dt = I ext + glEl (cid:104) gtot (cid:105) + Erev i i i gnoise (cid:104) gtot (cid:105) − u , (6) Cm duk dt = gl(El − uk) + I , (3) with capacitance Cm, membrane potential uk, leak po- tential El, leak conductance gl and input current I. The spiking condition is deterministic as well: when uk crosses a threshold ϑ from below, a spike is emitted and uk is reset to  for a refractory period τref (see Fig. 1B bottom). For conductance-based synapses, the synaptic input current to a neuron is typically modelled as dI syn k dt = − I syn k τsyn + (cid:88) (cid:88) syn i spk s with the synaptic time constant τsyn, the synaptic weight wki and the reversal potential of the ith synapse Erev . i The aim of this letter is to demonstrate how a network of deterministic neurons in a biologically plausible spik- ing noisy environment can quantitatively reproduce the stochastic dynamics required for sampling from a well- defined distribution p(z1, . . . , zK) and perform inference given observations. We start by calculating the dynam- ics of a single LIF neuron in a spiking noisy environ- ment and derive its activation function by describing the spike response as a first passage time (FPT) problem. This establishes an equivalence to the abstract, inher- ently stochastic units (2). On the network level, we show how biologically realistic conductance-based synapses (4) approximate the interaction for sampling from a well- defined target distribution. We complement our study with a demonstration of probabilistic inference by im- plementing the posterior of a small graphical model for handwritten digit recognition in a recurrent network of LIF neurons. Deterministic neurons in a noisy environment k + I ext obey eqn. (4), the current I ext k . While the synaptic currents I rec The total input current Ik to a neuron can be for- mally partitioned into recurrent synaptic input, diffuse synaptic noise and additional external currents: Ik = k + I noise I rec and I noise captures additional k current stimuli. We start by considering a single neuron that receives diffuse synaptic noise I noise in the form of random spikes from its surrounding. The capacity of re- current networks to produce such noise has been shown in [11]. Throughout the following analysis of individual neurons we omit the index k and set I rec = 0. k k k When a conductance-based LIF neuron receives strong synaptic stimulation, it enters a so-called high- conductance state (HCS, [12]), characterized by acceler- ated membrane dynamics. It is therefore convenient to time effective replaced by a smaller where the membrane time constant τm = Cm/gl is constant τeff = Cm/gtot, with the total conductance gtot subsum- ing both leakage and synaptic conductances. In a HCS, τeff governs the decay towards an effective leak poten- i + I ext)/gtot, where gnoise represents the total conductance at the ith synapse. In tial ueff = (glEl +(cid:80) a high input rate regime, (cid:112)Var(gtot)/(cid:104) gtot (cid:105) → 0 and i gnoise Erev i i the equation governing the membrane potential can be written as (cid:80) with (cid:104)·(cid:105) denoting the mean. In a first approximation, τeff can be considered very small in the HCS, resulting in u ≈ ueff , with the effective potential ueff simply being a linear transformation of the synaptic noise input. Using methods similar to [13], it can be shown that, if stimulated by a large number of uncorrelated spike sources, the synaptic current I noise – and therefore, also ueff – can be described as an Ornstein-Uhlenbeck (OU) process: du(t) = θ · (µ − u(t)) + Σ · dW (t) , (7) (8) (9) with parameters θ = τsyn I ext + glEl +(cid:80) (cid:88) (cid:2)wnoise νi i µ = Σ 2 = Erev i τsyn i i νiwnoise (cid:104) gtot (cid:105) (Erev i − µ)(cid:3)2 τsyn /(cid:104) gtot (cid:105) . (10) i where νi represents the input rate at the ith noise synapse and wi its weight. The activation function as an FPT problem The inherently stochastic neuron model (2) leads to a logistic activation function for constant potential v: p(z = 1) = σ(v) := [1 + exp(−v)] −1 . (11) In the following, we derive the activation function of the deterministic LIF neuron in a spiking noisy environment. Similarly to the abstract model [7], we define the refrac- tory state of a neuron as z = 1. An example of membrane potential dynamics with re- set is shown in Fig. 2A. Two modes of firing can be ob- served: the "bursting" mode, where the effective mem- brane potential after the refractory period is still above threshold, and the freely evolving mode, where the neu- ron does not spike again immediately after the refractory period. Denoting the relative occurrence of burst lengths n by Pn and the average duration of the freely evolving mode that follows an n-spike-burst by Tn, we can identify the following relation: (cid:80) (cid:80) n Pn · n · τon n Pn · (n · τon + Tn) p(z = 1) = . (12) Given the parameters of the associated OU process, we can derive a recursive expression for Pn and Tn, thereby ultimately allowing the calculation of p(z = 1): Pn = p(un < ϑ, un−1 ≥ ϑ, . . . , u1 ≥ ϑu0 = ϑ) (13) dun−1p(un−1un−1 ≥ ϑ) (cid:33)(cid:90) ∞ Pi ϑ (cid:35) dunp(unun−1) , dun−1p(un−1un−1 ≥ ϑ) = Tn = i=1 (cid:32) 1 − n−1(cid:88) (cid:34)(cid:90) ϑ (cid:90) ∞ (cid:34)(cid:90) ϑ −∞ ϑ −∞ dunp(unun < ϑ, un−1)(cid:104) T (ϑ, un)(cid:105) (cid:35) (14) . Fig. 2B displays an intuitive picture of the integrals in (13) and (14). The transfer function p(unun−1) is the Green's function of the OU process for t = τref : (cid:20) (un−(un−1−µ) exp(−θτref )−µ)2 (cid:21) − θ Σ 2 p(unun−1) = C e with the normalization C = (cid:112)θ/πΣ 2(1 − e−2θτref ). 1−exp(−2θτref ) , (15) T (ub, ua) denotes the time the membrane needs to reach ub starting from ua. This FPT problem has been ex- tensively discussed in literature and its moments can be given in closed form [14]. To improve the prediction of the activation function, we further take into account small, but finite τeff , in which case the membrane potential no longer directly fol- lows the input current, but is a low-pass-filtered version thereof. By using an expansion in (cid:112)τeff /τsyn, a first- order correction to the FPT can be calculated [15]: (cid:104) T (ϑ, u)(cid:105) = τsyn √ π dx exp(x2)[erf(x) + 1] with the effective threshold ϑeff ≈ ϑ−ζ(cid:0) 1 u−µ σ 2 (cid:1)(cid:113) τeff 2τsyn , (16) , where ζ denotes the Riemann Zeta function. A comparison of the predicted p(z = 1) with results from a numerical simulation is shown in Fig. 2C. Here, the average effective potential ¯u = (cid:104) ueff (cid:105) was established through an external current I ext. For the translation from the LIF domain to σ(cid:90) ϑeff −µ 3 FIG. 2. (A) Membrane potential u(t) (blue) and resulting spike activity (black) of a LIF neuron in a spiking noisy en- vironment. (B) Overlay of u (blue) and ueff (red). Due to the small τeff in the HCS, the two curves are nearly identical when the neuron is not refractory. At the end of each refrac- tory state (corresponding to z = 1, grey), the predicted prob- ability distribution for ueff is plotted in pink. The normalized subthreshold area (dark pink) is used for the propagation in (13). (C) Theoretical prediction (red) vs. simulation results (blue); errors are smaller than the symbol size. A logistic function σ(¯u) (green) has been fitted to the prediction. the abstract stochastic model (2) we identify ¯u − ¯u0 α , v = (17) ¯u0 the value ¯u for which where denotes p(z = 1) = σ(0) = 1 2 and α represents a scaling fac- tor between the two domains. We conclude that a single LIF neuron in a spiking noisy environment can closely reproduce the activation function (11). of Sampling via recurrent networks of LIF neurons with conductance-based synapses k We next connect the neurons to a recurrent network. In addition to noise stimuli, a LIF neuron receives synap- tic currents I rec from other neurons in the ensemble. Synaptic interaction introduces correlations among the spike response of different neurons, i.e. the random vari- ables z1, . . . , zK are not independent. For certain con- nectivity structures, it is possible to specify a target dis- tribution [7, 16] for the network states z(t) that occur under the network dynamics . In the following, we use the emulation of Boltzmann machines as an example case. The joint distribution reads: pB(z) = 1 Z exp zTW z + zTb , (18) (cid:18) 1 2 (cid:19) K(cid:88) (cid:17) z(cid:48) exp (cid:80) (cid:16) 1 2 z(cid:48)TW z(cid:48) + z(cid:48)Tb where W is a symmetric zero-diagonal weight ma- trix, b is a bias vector, zTW z and zTb denote bi- linear forms over W and 1 respectively, and Z = is the partition function that ensures correct normalization. This probabilistic model underlies state-of-the-art machine learning algo- rithms for image [17] and speech recognition [18]. For the abstract neuron model (2), it had been proven [7] that a membrane potential of the form vk = bk + Wkj zj (19) j=1 leads to the desired target distribution (18) of network for t → ∞. This finding uses the fact states z(t) that individual neurons can sample from the condition- als p(zk = 1 z\k) = σ(vk), with z\k = {zj j (cid:54)= k}, in a Gibbs sampling inspired updating scheme. As shown above, LIF neurons in a spiking noisy envi- ronment closely approximate this logistic activation func- tion if the synaptic currents I rec shift the mean mem- brane potential ¯uk according to the linear interaction (19). Using the linear transformation (17) between vk and ¯uk, and estimating the impact of a pre-synaptic spike on the post-synaptic neuron through conductance-based synapses of weight wkj, we arrive at the following pa- rameter translation between the abstract and the LIF domain: bk =(¯ub Wkj = k − ¯u0 wkj 1 k)/α (cid:16) (cid:17) kj − µ Erev (cid:0)e−1 − 1(cid:1) − τeff 1 − τsyn τeff τsyn αCm (cid:104) (20) (cid:16) e (cid:17)(cid:105) − τsyn τeff − 1 , (21) 4 FIG. 3. (A) Spike pattern of a recurrent network of LIF neurons during sampling from a randomly generated Boltz- mann machine. (B) Sampled distribution pN(z) of network states (blue bars) and analytically calculated target distribu- tion pB(z) (in red). (C) Kullback-Leibler divergence between the sampled distribution pN and the target distribution pB as a function of integration time T for 10 trials (thin lines). The red dotted line shows convergence for a network of intrinsi- cally stochastic (theoretically optimal) neurons that is guar- anteed to converge to pB for T → ∞. (D) Kullback-Leibler divergence DKL (pN pB) when sampling for T = 103 s from 100 different randomly generated target distributions. where ¯ub k is the mean free potential ¯uk in Fig. 2C that establishes p(zk = 1 z\k = 0) = σ(bk), and Erev kj denotes the reversal potential for synapse wkj. The idea behind (21) is to match the integrals of individual postsynaptic potentials (PSPs) on vk and ¯uk. We furthermore employ short-term synaptic depression to approximate the the- oretically optimal rectangular PSP shape also in case of consecutive spikes (bursts). The sampling process with networks of LIF neurons was examined in computer simulations. Fig. 3A shows the spike pattern of a recurrent network of K = 5 LIF neurons that sample from a randomly generated Boltz- mann machine. The parameters bk and Wkj were drawn from the interval [−0.6, 0.6]. The thus defined target distribution pB(z) is approximated by the distribution pN (z) of network states when the spike pattern is inter- preted as samples z(t) by convolution with a τon = 10 ms box kernel. Fig. 3B shows the average network distri- bution pN(z) (blue bars) after T = 10 s simulation time alongside the target values pB(z) (red lines) calculated from (18). Sampled probabilities pN(z) depict the mean over 10 independent simulation runs, errorbars reflect stochastic variations between individual runs. The cho- sen integration time T = 10 s displays a conservative es- timate of the maximum duration a neuronal ensemble experiences stable stimulus conditions in a behaving or- ganism and can thus be expected to sample from a sta- tionary distribution. We find that the recurrent network of LIF neurons accurately encodes the target distribution over several orders of magnitude, within the precision im- posed by the sample-based representation. Fig. 3C shows how the network distribution becomes increasingly more reliable as more samples are consid- ered. After few samples, the network has generated a coarse approximation of pB(z) that could serve as an "ed- ucated guess" in online computation tasks. For simula- tion times T well beyond biologically relevant timescales, systematic errors in pN (z) become apparent: The KL di- vergence saturates on a non-zero value, while the (the- oretically ideal) abstract model [7] further converges to- wards the target distribution (red dotted line). Fig. 3D shows that the sampling quality holds for a variety of similarly generated target distributions. Demonstration of probabilistic inference We conclude our investigation of sampling in recurrent networks of LIF neurons with an example of Bayesian inference based on incomplete observations. A fully con- nected Boltzmann machine of K = 144 neurons, aligned on a 12x12 grid, was trained as an associative network [19] to store three prototypic patterns, namely hand- written digits 0, 3 and 4. In the network, each cell is assigned to one pixel of the image. Statistical correla- tions between the pixels as well as their mean intensities were encoded in the weights Wkj and biases bk of the corresponding joint distribution p(z). This distribution reflects "prior knowledge" imprinted in the network. The probabilistic model is augmented by adding real- valued input channels for each pixel. Like the network variables zk, input channels are associated with random variables yk ∈ R, 1 ≤ k ≤ K. The resulting graphical model is sketched in Fig. 4A and entails the following structure for a full probabilistic model: p(y, z) = p(z) · K(cid:89) p(yk zk) . (22) k=1 The full model p(y, z) connects the network variables zk of the prior to inputs yk by means of the likelihood p(yk zk). We have chosen a Gaussian likelihood with unit variance (see Fig. 4B): p(yk zk) = N (yk; µ = zk − 1 2 , σ2 = 1) . (23) The task for the network is to implement the posterior distribution that follows from Bayes' rule: p(z y) ∝ p(z) · p(y z). The posterior combines two sources of information: The likelihood p(yk zk) tends to align the network state with the observation, i.e. zk = 1 for yk > 0, while the prior p(z) reconciles the observations with knowledge on consistent activation patterns z. In this way, the posterior p(z y) evaluates all possible outcomes z y simultaneously by assigning a belief to each of them, and thus captures the model's (un-)certainty about dif- ferent solutions. A short derivation shows that the poste- rior p(z y) is a Boltzmann machine for any input y, thus being compatible with the sampling dynamics of spiking networks. More specifically, we obtain the following ab- stract membrane potential: vk = bk + yk + Wkj zj . (24) j In the LIF domain, the sum bk + yk is equivalent to an effective bias (20) and corresponds to an external cur- rent I ext k that shifts ¯uk appropriately. Thus, k = I b k + I y (cid:88) 5 k + I b k + I y k . In case of I y a network neuron receives synaptic input from recurrent connections and noise sources, as well as an external cur- k + I noise rent, i.e., Ik = I rec k = 0 ∀k, the network samples from the prior distribution p(z) = p(z y = 0). A two- dimensional projection of network states z(t) ∼ p(z) is shown in Fig. 4C. The sampled distribution exhibits three distinct peaks that correspond to the three hand- written digits stored in the recurrent weight matrix. A closer look at the network trajectory reveals how the system stays in one mode for some duration, traverses the state space and then samples from a different mode of the distribution. These dynamics also reflect in the marginals of the network variables under a 20 ms box k dt(cid:48), shown in the color filter, ¯zk(t) = 1 maps. (cid:82) t t−20 ms z(t(cid:48)) 20 ms A typical scenario of stochastic inference on incomplete observations is shown in Fig. 4D. Four input channels, located at the center of the grid, were picked to inject positive currents I y k > 0 to the network while all other inputs remained uninformative. Positive currents I y k en- code positive values of the respective input pixels yk and were chosen such that the observation appeared incom- patible with the digit 0, and remained ambiguous with re- spect to digits 3 and 4. Accordingly, in the posterior dis- tribution p(z y) the 0-mode is significantly suppressed while uncertainty about the provided cue is expressed by two distinct modes in the 3 and 4 directions. Discussion We have shown how recurrent networks of determin- istic neurons in a spiking noisy environment can per- form probabilistic inference through sampling from a well-defined posterior distribution. Our approach builds on theoretical work by Buesing et al. [7] and extends Bayesian spiking network implementations to determin- istic neuron models widely used in computational neu- roscience. For the analytical derivation and the com- puter simulations we have employed leaky integrate-and- fire neurons with conductance-based synapses. However, the analysis can be readily transferred to other neuron models [20]. The essential diffusion approximation relies on high-frequency spiking inputs that could be provided by the surrounding network and lead to strong synaptic conductances and fast membrane dynamics. Thereby, our derivation identifies a potential functional role of bi- ologically observed high-conductance states within a nor- mative framework of brain computation. For mathematical tractability, simplifying modeling assumptions had to be made. The neuron model only uses an absolute refractory time τref , which matches the activation time constant τon, and neglects any ad- ditional gradual recovery effects after a spike. On the network level, we have assumed statistically indepen- 6 FIG. 4. (A) Graphical model used for the probabilistic inference task. The network implements a Markov random field over latent variables z. Observables yk are conditionally independent given the network state. (B) A mixture-of-Gaussians likelihood model provides input to the sampling neurons. (C) Two dimensional projection of sampled states z(t) when sampling from the prior p(z). The network preferentially spends time in modes close to the stored hand-written digits 0, 3, 4. Solid line: network trajectory over 200 ms. Color maps: Marginals ¯zk(t) averaged over 20 ms. The time arrow covers the duration of the red trajectory and consecutive snapshots are 20 ms apart. (D) As in (C) when sampling from the posterior p(z y) with incomplete observations y. The provided input y is incompatible with digit 0 and ambiguous with respect to digits 3 and 4. dent noise sources and instantaneous axonal transmis- sion. Furthermore, post-synaptic potentials mediated through conductance-based synapses differ from the the- oretically optimal rectangular shape and can lead to de- viations from the target distribution outside of the high noise regime [21]. However, computer simulations indi- cate that in most scenarios the above approximations are not critical. Beyond neuroscience, the ability to perform proba- bilistic inference with deterministic neurons displays a promising computing paradigm for neuromorphic hard- ware systems. Originally designed as neuroscientific modeling tools, these systems typically implement a physical model of integrate-and-fire neurons [22, 23], which renders the application of the proposed networks straightforward. In particular, the distributed nature of the sampling algorithm allows to exploit the inher- ent parallelism of neuromorphic architectures, fostering an application of neuromorphic hardware to online data evaluation and robotics. ∗ The first two authors contributed equally to this work. We thank W. Maass for his essential support and Ste- fan Habenschuss for helpful comments. This research was supported by EU grants #269921 (BrainScaleS), #237955 (FACETS-ITN), the Austrian Science Fund FWF #I753-N23 (PNEUMA) and the Manfred Stark Foundation. [5] P. Berkes, G. Orb´an, M. Lengyel, and J. Fiser, Science 331, 83 (2011). [6] S. Deneve, Neural computation 20, 91 (2008). [7] L. Buesing, J. Bill, B. Nessler, and W. Maass, PLoS computational biology 7, e1002211 (2011). [8] R. P. Rao, in Advances in neural information processing systems (2004) pp. 1113–1120. [9] P. O. Hoyer and A. Hyvarinen, Advances in neural infor- mation processing systems , 293 (2003). [10] Z. F. Mainen and T. J. Sejnowski, Science 268, 1503 (1995). [11] N. Brunel, Journal of computational neuroscience 8, 183 (2000). [12] A. Destexhe, M. Rudolph, and D. Pare, Nature Reviews Neuroscience 4, 739 (2003). [13] L. M. Ricciardi and L. Sacerdote, Biological Cybernetics 35, 1 (1979). [14] L. M. Ricciardi and S. Sato, Journal of Applied Proba- bility 25, 43 (1988). [15] N. Brunel and S. Sergi, Journal of Theoretical Biology 195, 87 (1998). [16] D. Pecevski, L. Buesing, and W. Maass, PLoS compu- tational biology 7, e1002294 (2011). [17] R. Salakhutdinov and G. E. Hinton, in International Conference on Artificial Intelligence and Statistics (2009) pp. 448–455. [18] A.-R. Mohamed, G. E. Dahl, and G. Hinton, Audio, Speech, and Language Processing, IEEE Transactions on 20, 14 (2012). [19] G. Hinton, P. Dayan, B. Frey, and R. Neal, Science 268, 1158 (1995). [20] R. Brette and W. Gerstner, Journal of neurophysiology 94, 3637 (2005). [21] W. Gerstner and J. L. van Hemmen, Network: Compu- [1] K. P. Kording and D. M. Wolpert, Nature 427, 244 tation in Neural Systems 3, 139 (1992). (2004). [22] S. Mitra, S. Fusi, and G. Indiveri, Biomedical Circuits [2] J. Fiser, P. Berkes, G. Orb´an, and M. Lengyel, Trends and Systems, IEEE Transactions on 3, 32 (2009). in cognitive sciences 14, 119 (2010). [3] K. Friston, J. Mattout, and J. Kilner, Biological cyber- netics 104, 137 (2011). [4] T. Yang and M. N. Shadlen, Nature 447, 1075 (2007). [23] D. Bruderle, M. A. Petrovici, B. Vogginger, M. Ehrlich, T. Pfeil, S. Millner, A. Grubl, K. Wendt, E. Muller, M.- O. Schwartz, and et al., Biological Cybernetics 104, 263 (2011).
1710.07613
1
1710
2017-10-15T07:24:12
Quantum field inspired model of decision making: Asymptotic stabilization of belief state via interaction with surrounding mental environment
[ "q-bio.NC" ]
This paper is devoted to justification of quantum-like models of the process of decision making based on the theory of open quantum systems, i.e. decision making is considered as decoherence. This process is modeled as interaction of a decision maker, Alice, with a mental (information) environment ${\cal R}$ surrounding her. Such an interaction generates "dissipation of uncertainty" from Alice's belief-state $\rho(t)$ into ${\cal R}$ and asymptotic stabilization of $\rho(t)$ to a steady belief-state. The latter is treated as the decision state. Mathematically the problem under study is about finding constraints on ${\cal R}$ guaranteeing such stabilization. We found a partial solution of this problem (in the form of sufficient conditions). We present the corresponding decision making analysis for one class of mental environments, the so-called "almost homogeneous environments", with the illustrative examples: a) behavior of electorate interacting with the mass-media "reservoir", b) consumers' persuasion. We also comment on other classes of mental environments.
q-bio.NC
q-bio
Quantum field inspired model of decision making: Asymptotic stabilization of belief state via interaction with surrounding mental environment Fabio Bagarello 1 Irina Basieva 2 Andrei Khrennikov 3 Abstract This paper is devoted to justification of quantum-like models of the process of decision making based on the theory of open quantum systems, i.e. decision making is consid- ered as decoherence. This process is modeled as interaction of a decision maker, Alice, with a mental (information) environment R surrounding her. Such an interaction gen- erates "dissipation of uncertainty" from Alice's belief-state ρ(t) into R and asymptotic stabilization of ρ(t) to a steady belief-state. The latter is treated as the decision state. Mathematically the problem under study is about finding constraints on R guaranteeing 1 Dipartimento di Energia, Ingegneria dell'Informazione e Modelli Matematici, Scuola Politecnica, Universit`a di Palermo, I-90128 Palermo, INFN, Sezione di Napoli, ITALY and Department of Mathematics and Applied Mathematics, Cape Town, South Africa e-mail: [email protected] Home page: www1.unipa.it/fabio.bagarello 2 Department of Psychology, City University, London, UK e-mail: [email protected] 3 International Center for Mathematical Modeling, in Physics and Cognitive Science Linnaeus University, Vaxjo, Sweden National Research University of Information Technologies, Mechanics and Optics (ITMO) St. Petersburg, Russia e-mail: [email protected] such stabilization. We found a partial solution of this problem (in the form of suffi- cient conditions). We present the corresponding decision making analysis for one class of mental environments, the so-called "almost homogeneous environments", with the illus- trative examples: a) behavior of electorate interacting with the mass-media "reservoir"; b) consumers' persuasion. We also comment on other classes of mental environments. keywords: decision making; quantum-like model; mental (information) environment; open quantum systems; dissipation of uncertainty; voters' behavior; consumers' persuasion 1 Introduction The recent years were characterized by explosion of interest in applications of the mathematical formalism of quantum theory to studies in cognition, decision making, psychology, economics, finance, and biology, see, e.g., the monographs [1]-[6] and a few representative papers [7]-[33] (the first steps in this direction were done long time ago, see, e.g., [34]). The approach explored in such mathematical modeling is known as quantum-like. In this approach an agent (human, animal, or even cell) is considered as a black box processing infor- mation in accordance with the laws of quantum information and probability theories. Thus the quantum-like modeling is basically quantum informational modeling (although this character- istic feature is typically not emphasized, cf., however, with [35]).1 Quantum-like models have to be sharply distinguished from genuinely quantum physical models of cognition which are based on consideration of quantum physical processes in the brain, cf. with R. Penrose [38] and S. Hameroff [39]. Although the quantum physical models have been criticized for mismatching between the temperature and space-times scales of the quantum physical processes and neuronal processing in the brain, see especially Tegmark [40], they cannot be rejected completely and one may expect that quantum-like models of cognition will be (soon or later) coupled with real physical processes in the brain, see [41]-[47] for some steps in this direction. The quantum-like approach generated a variety of models of cognition and decision mak- ing. In the simplest model [34], [1], the mental state (the belief state) of an agent, Alice, is represented as a quantum state ψ and questions or tasks as quantum observables (Hermitian operators). Answers to the questions are given with probabilities as determined by Born's rule. 1See, e.g. D' Ariano [36] and Plotnitsky [37] for the information approach to quantum mechanics. 2 For Hermitian operator A with the purely discrete spectrum, Born's rule can be written as p(A = αk) = kPαkψk2 = hPαk ψ, ψi, (1.1) where αk is an eigenvalue of A and Pαk is the projector on the eigenspace corresponding to this eigenvalue. This model does not describe dynamics of the belief state in the process of decision making. Consideration of dynamics was introduced in the works of Khrennikov [7, 8], and Pothos and Busemeyer [10]. In their dynamical model as in the previous models, an observable A corre- sponding to a question (task) faced by Alice is represented as a Hermitian operator. Then Hamiltonian H generating unitary dynamics ψ(t) = U(t)ψ0, U(t) = e−itH (1.2) of the initial belief state ψ0 is introduced, and Alice's decision is represented as measurement of the observable A at some instant of time. The authors of [10] presented cognitive arguments supported by experimental studies to determine the instant tm of measurement. Here the probability of a particular answer is also determined by the Born rule, but applied to belief state ψ(tm). Of course, this is an important issue, since different values ot tm can give rise to completely different results. In spite of the partial progress in determination of tm in the article of Pothos and Busemeyer [10], this complex problem cannot be considered completely solved. Moreover, in solving this problem Pothos and Busemeyer had to go beyond the quantum theory and to appeal to psychological theoretical and experimental studies. It would be attractive to solve this problem entirely in the quantum framework. We remark that the Schrodinger equation describes the dynamics of an isolated system. In the presence of an environment, the dynamics of the system is non-unitary. Approximately (un- der some sufficiently natural conditions) it is described by the Gorini-Kossakowski-Sudarshan- Lindblad (GKSL) equation (often called simply the Lindblad equation), the simplest version of the quantum master equation. One of the main distinguishing features of such dynamics is that it does not preserve the pure state structure: it (immediately) transforms a pure initial state ψ0 into a mixed state given by the density operator: ρ(t) = U(t)ρ0, U(t) = e−itL, (1.3) where L is the generator of the GKSL-evolution and ρ0 = ψ0ihψ0. This dynamics is in general non-unitary. This equation describes the process of the system adaptation to the surrounding environment. This is the complex dynamical process combining the internal state dynamics 3 of the system with adaptation to signals received from the environment. If the dynamics is discrete with respect to time, then it can be represented as a chain of unitary evolutions and (generalized) quantum Bayesian updates.2 In this paper we cannot discuss this interesting issue in more detail, see [16] for detailed consideration of a two dimensional example with application to the evolution theory. In fact, GKSL-dynamics does not contradict the Schrodinger equation structure of the quan- tum evolution. Let us denote the system under study by S and the surrounding environment by R ("reservoir" for S.) Suppose that initial state of the compound system S + R is pure and separable. Dynamics in the state space of S + R is still unitary and given by a Hamiltonian for S + R. The main distinguishing feature of this unitary dynamics is that (in the presence of interaction between S and R) it induces entanglement and the state of the compound system becomes not more separable. Hamiltonians for the composite system S + R are very complex, since they includes, in general, an infinite number of degrees of freedom of R. Typically it is im- possible to solve the Schrodinger equation for the state of composite system S + R. (Although in some special cases, as those considered in this paper and others discussed in [4] analytic solutions can be found.) Therefore, most studies are restricted to the dynamics of the state ρ(t) of S alone, which is described (approximately) by the GKSL-equation. But even if one were able to solve the Schrodinger equation for S + R, the solution would be a very complex infinite-dimensional state vector. Since we are interested in behavior of S, we would then take the trace with respect to all degrees of freedom of R and obtain the state of S. A simple math- ematical theorem implies that in presence of entanglement this trace-state cannot be pure, i.e., the state is descirbed by density operator. Its dynamics under the GKSL-equation is known as decoherence: decreasing of state's purity (or coherence) in the process of interaction with an environment. Since consideration of an isolated cognitive system is even a higher degree idealization than consideration of an isolated physical system, it is natural to modify the dynamical scheme of de- cision making based on unitary Schrodinger dynamics [7, 8, 10] and consider general dynamics of the belief-state, either by using the approximative GKSL-dynamics or by tracing the state of 2As was quickly understood in quantum physics, the Luders projection postulate describes only one very special class of the quantum state updates resulting from measurements. We remark that already von Neumann accepted applicability of this straightforward form of the state update only for observables with non-degenerate spectra. Generally, in the case of an observable with degenerate spectrum, a pure pre-measurement state can be transferred into a mixed state [48]. Later these considerations of von Neumann were elaborated in the form of the theory of quantum instruments, see [49] for non-physicist friendly presentation. The most consistent justification of this theory is obtained in the framework of the theory of open quantum systems. 4 the compound system. Roughly speaking there is no choice: either one has to ignore the pres- ence of environment or consider non-unitary dynamics, e.g., (1.3). Of course, such non-unitary dynamical model of decision making is much more mathematically complicated. However, it has one very important advantage. Here the aforementioned problem of determination of tm is solved automatically, although not straightforwardly. For a natural class of quantum master equations, in the limit t → ∞ the system's state ρ(t) approaches a steady state ρout. Diagonal elements of this operator (with respect to the "pointer basis" corresponding to the observable A) under measurement give the probabilities of measurement outputs. The model of decision making as decoherence was represented in a series of papers of Asano et al., see, e.g., [51], [52] [13], [6]. The model is purely informational. Both a "quantum-like system", Alice, and her mental (or information) environment ("bath", reservoir") are represented by quantum states, ψ and φ. (It can be assumed that initially these states are pure.) We are not interested in their physical or neurophysiologic realizations. Asano et al. [51] applied the theory of open quantum systems and the GKSL-equation to model experimental data collected in decision making experiments. The initial belief state is typically represented as a pure state of complete uncertainty, ψ = (0i + 1i)/√2, (1.4) where ii, i = 0, 1, are eigenstates of the dichotomous observable A representing a "yes"/"no" question. A proper decision making dynamics should asymptotically (t → ∞) drive the belief state of Alice ρ(t) to the output (mixed) state represented by the density operator ρout. Its diagonal elements in the basis of eigenvectors of the operator A give probabilities of possible decisions. We remark that the open quantum systems approach to decision making can be considered as a possible realization of the contextual treatment of cognition, cf. Khrennikov [1], [2] and Dzhafarov et. al. [12], [17]. The surrounding mental environment represents a measurement context for decision making. Of course, Alice cannot wait for t → ∞ to make a decision. And, as we know in physics, a quantum system relaxes to a steady state very quickly. Here we have to point to difference between a mathematical model and its applications to real phenomena. The notion of limit is a mathematical abstraction. Of course, it is not applicable to real physical or cognitive systems. In practice, the limit procedure encodes the process of approaching some quantity. The existence of a limit for the density operator ρ(t) guarantees that its fluctuations decrease. If the magnitude of fluctuations becomes smaller than some ǫ > 0, then such ρ(tǫ) can be selected 5 as a good approximation of the steady state (the latter is also a mathematical abstraction, in real physical processes it is never approached exactly). This tǫ plays the role of tm from the unitary dynamical model of decision making. Moreover, more complicated considerations based on the theory of decoherence demonstrate that this instant of time tǫ can be identified with the relaxation time T. The latter is determined by the structure of interactions between a system and its environment. Typically in quantum physics the time interval [0, T ] is very short and a system relaxes very quickly to its steady state. Since such considerations would make the paper even more complex mathematically, we shall not proceed in this direction. (See, e.g., [53] for derivation of an analytical expression for the relaxation time as a function of the heat-bath and interaction parameters.) The main message of these considerations to a reader is that limt→∞ is just an abstraction (at t = ∞ all fluctuations disappear completely). In reality a system relaxes very quickly to its steady state. The main problem of the quantum(-like) decision theory is to construct an operator repre- sentation of Hamiltonians and "Lindbladians" (the latter represents interactions of Alice with her mental reservoir). In contrast to quantum physics, there is no analog of the classical phase formalism for cognition and decision making, i.e., we cannot use the quantization procedure to transform functions on the classical phase space into operators - the procedure of Schrodinger's quantization is not applicable. In quantum theory one can also use the quantization procedure based on the algebra of operators of creation and annihilation. It is especially useful for second quantization (quantum field theory), see appendix 1 for brief presentation of the basics. This sort of quantization can be successfully applied to quantum-like modeling. Consider again a dichotomous question A, with the values "no"/ "yes". Creation operator a⋆ creates "yes" from "no", annihilation operator a transforms "yes" to "no". We can compound Hamiltonians and observables with the aid of creation and annihilation operators (similarly, e.g., to quantum optics) . This approach to construction of operators for problems of decision making was pioneered by Bagarello [4] and coauthors who applied it, see, e.g., [28], [26], [27], to a variety of problems. Bagarello et al. were interested in time dynamics of averages; in particular, probabilities were treated statistically. In series of works [28], [26], [27], a system was interpreted as an agent and the reservoir had a purely information interpretation, see especially [54] for a discussion. The same approach based on creation-annihilation operators can be explored to model decision making at individual level (with the subjective interpretation of probabilities). Here one can explore the scheme which was approved in the works of Asano et al. [6]: consideration of the asymptotic dynamics of the belief state and using the asymptotically output state (for 6 t → ∞) as the basis of decision making. The first step in this direction was done in paper [51]. The aforementioned decision making scheme (based on asymptotic stabilization of the mental (belief) state of an agent) generates interest to study the asymptotic dynamics of the system interacting with a reservoir. This is a nontrivial mathematical problem and its partial solution (for specially designed information reservoirs) is presented in this paper. In section 2 we discuss several aspects of the process of decision making for a dichotomous question A through belief-state stabilization. Then, in section 3, the mathematical problem is discussed in more details. More explicitly, in section 3.1 we prove stabilization of the belief state in the case of the "almost homogeneous mental environment" of Alice. We start with the case of a question' A having an infinite number of possible outcomes labeled as na = 0, 1, ..., n, ... . Then in section 3.3 we consider the case of a "dichotomous question" A having two possible outcomes, na = 0, 1. In section 3.2 we generalize our theory to the case of environments having a more complex structure. Section 4 contains our conclusions, while appendices 1 and 2 are devoted to few introductory remarks on canonical anti-commutation relations (CAR) and some extra mathematical details. 2 Decision making as decoherence Consider the case of two dichotomous questions posed to Alice, A = 0, 1 ("no", "yes") and the mental environment R similarly composed of dichotomous degrees of freedom. In section 3.3 this situation will be modeled by using operators a, a† (Alice's operators) and b, b† (R's operators) satisfying the CAR: {a, a†} = I,{b, b†} = I,{a, b} = 0, (2.1) where the anti-commutator of two operators x, y is defined as {x, y} = xy + yx. Moreover, a2 = b2 = 0. In quantum field theory, see appendix 1, these operators are known as the operators of creation and annihilation.3 The operators related to R can depend on some parameter k (discrete or continuous) representing degrees of freedom of R, b = b(k), b† = b†(k). Typically environment R has a huge number of degrees of freedom and belongs to infinite-dimensional Hilbert state space K. In the case of dichotomous questions asked to Alice, it can be assumed that her state space H has the dimension of two, the qubit space of quantum information theory. In quantum field theory the terminology "creation-annihilation operators" has coupling to real physical systems. The operators represent the processes of creation-annihilation of quantum 3In physics these operators represent the processes of creation and annihilation of fermions, e.g., electrons. 7 particles (e.g., photons or electrons), see appendix 1. As was emphasized in introduction, our model is of the purely informational nature. Therefore the operators a, a† and b, b† have to be treated as the formal representation of the processes of "creation" and "annihilation" of information states.4 Consider, e.g., Alice's operators a, a†. Let φ0 = 0i, φ1 = 1i be the orthonormal basis in Alice's state space H corresponding to the answers "no"/"yes" to the question A. As a consequence of the CAR, the operators a, a† act on these basis vectors in the following way: a†0i = 1i, a†1i = 0; a1i = 0i, a0i = 0. The operators a and a† modify Alice's attitude to selection of alternatives. In this framework these operators are considered as reflection operators [54]. We represent the question A as the number operator of Alice, na = a†a. The eigenvalues of this operator, na = 0, 1, correspond to the choices of Alice at t = 0. Therefore it is natural to name na the decision operator.5 Consider the simplest situation which can be modeled in this framework: Alice is asked a question A and she is surrounded by a population R whose members are asked the same question A. Alice interacts with this population R and she gets to know behavior of its members with respect to this question. Of course, she cannot "scan" R completely to know the concrete answers of people to A. She just gets to know a sort of average nb of answers to A. In the simplest case of a homogeneous population this average is a constant. In the general case average nb can non-trivially depend on the parameter k encoding various population clusters R(k) : nb = nb(k). Here Alice (through interaction with this population) obtains information about averages for the answers to A corresponding to different clusters R(k) of R. In principle, R needs not be combined of physical agents. As was emphasized in introduc- tion, our model is of the purely informational nature. The environment R can, for example, represent mass-media's image of the question under consideration: the image created in TV- debates and shows, web-blogs, newspapers. For example, A can be a referendum question: "vote for Brexit or against?" (or the recent USA-election question: "vote for Trump or not?"). By analysing (generally unconsciously) mass-media's opinions Alice estimates the average nb (or in general the averages nb(k)) and she makes her decision based on these averagess. In a more general situation the members of R express their attitude with respect to a variety of dichotomous questions and the A, the question asked to Alice, can be among them, but not necessarily. In the latter case Alice makes her decision based on behavior of the surrounding 4In principle, we can simply call these operators ladder operators as is done in formal mathematical theory. 5In quantum field theory the number operator has the meaning of the number of physical particles, see appendix 1. In our model it represents states indexed by numbers. 8 environment R (physical or informational) without even understanding that her decision is made through influence of R. Now we consider dynamics of decision making. In section 3 we shall use the Heisenberg picture, i.e., we are concerned with the dynamics of operators. We are interested in the dynamics of the reflection operators, a†(t), a(t), with initial conditions a†(0) = a†, a(0) = a. Based on the reflection operators dynamics we can find the the decision operator dynamics na(t) = a†(t)a(t) and, hence, its average na(t) :=< na(t) ⊗ I >=< a†(t)a(t) ⊗ I > (2.2) with respect to some initial state of the compound system, see section 3 for details. In the case of dichotomous questions the average na(t) has the straightforward probabilistic meaning. To see this, let us expand expression (2.2). Denote the state space of the compound system, Alice and her mental environment, by the symbol H⊗K and the operator of Heisenberg evolution of this system by Ut. For an arbitrary initial state h., .iR, formula (2.2) reads as na(t) =< Ut(a† ⊗ I)U † t Ut(a ⊗ I)U † t >R= TrUt(a†a ⊗ I)U† tR. By using the cyclic property of the trace we get: tRUt = TrR(t)(a†a ⊗ I) =< a†a ⊗ I >R(t), na(t) = Tr(a†a ⊗ I)U† where R(t) = U † t RUt represents the state dynamics (so we transferred the operator dynamics into the the state dynamics, from the Heisenberg picture to the Schodinger picture).6 Consider now the partial trace of R(t) with respect to the environmental degrees of freedom, ρ(t) = TrKR(t). It represents the dynamics of Alice's belief-state in the process of her interaction with the environment. Consider the average of the decision operator na = a†a = a†(0)a(0) with respect to this belief-state: < na >ρ(t)= Trρ(t)na. This average is directly related to probability. Since the decision operator na has two eigenval- ues, na = 0, 1, the average coincides with the probability: Pt(A = 1) =< na >ρ(t)=< ρ(t)φ1, φ1 > . (2.3) 6The Schrodinger picture providing the belief-state dynamics is basic for the probabilistic interpretation. However, equations in the Heisenberg picture are easier for analytic treatment. Therefore in section 3 we shall work in the Heisenberg picture. 9 We recall that here the question A is represented by the operator na. The eigenvector φ1 corresponds to the answer "yes" to this question. Pt(A = 1) is the subjective probability in favor of this answer which Alice assigns at the moment of time t. We remark that this assignment can be done unconsciously. Alice will consciously report only her final decision, see considerations below. Now we make a very general remark about the partial trace. For any operator M acting in H, the following equality holds: TrρM = TrR(M ⊗ I), where ρ = TrKR. Therefore the average na(t) equals to the average of Alice's decision operator: na(t) =< na >ρ(t) . (2.4) In section 3 we obtain the averages of the decision operator na(t) for its eigenstates φna (for operators satisfying CAR, we have: na = 0, 1). Thus initially Alice has a definite belief about possible answers to the question A. In section 3.1 we consider the case of the "almost homogeneous reservoir" - a mental (in- formation) environment R which is behaviorally almost homogeneous in the following sense (see section 3.1 for the mathematical formalization). The basic parameter characterizing R is deduced by < b†(k)b(q) >= nb(k)δ(q − k), the average for k-type agents. For an almost homo- geneous R, nb(k) is constant. In general, however, it can vary with k 7. It must be stressed that the reservoir considered in section 3.1 is not completely homogeneous. Its members are not all identical. They have different internal dynamical scales represented by the function ω(k) = ωbk which really depends on k, even if this dependence is relatively mild (ω(k) is linear in k, so that ω(k1) ≃ ω(k2) if k1 ≃ k2), see again section 3.1. For such environment R, the average na(t) of the decision operator na was found, see formula (3.7) section 3.1 (with the corresponding generalization to the CAR-case in section 3.3). This formula in combination with equalities (2.3), (2.4) gives the probabilistic dynamics: Pt(A = 1) = na(t) = na e− 2λ ω b 2 π t + nb(cid:18)1 − e 2 π 2λ ω b t(cid:19) = nb + (na − nb) e− 2λ ω 2 π b t, (2.5) where λ is the constant of interaction between Alice and R, see (3.1). As we will see in the 7In the simplest model all agents in R are asked the same question A as Alice and nb(k) is the average of their answers to A. In a more complex environment R each cluster R(k) of agents labeled by the parameter k is asked its own question A(k) and nb(k) is the average of their answers to A(k). 10 next section, this is the result of a reasonable choice of the Hamiltonian operator, where the interactions between Alice and her reservoir are included8. Consider, e.g., the case na = 1, e.g., initially Alice was for Brexit. Suppose that her surrounding environment is almost homogeneous9 and nb = 1/3, i.e., its members express sufficiently strong anti-Brexit attitude. Then Alice following the dynamics expressed by (2.5) would lose her inclination to vote for Brexit and finally she may accept the surrounding attitude to vote against Brexit. It is interesting to notice that the rapidity of the decision process is directly related to the strength of the interaction between Alice and her environment, and inversely connected to ωb. This is in agreement with the fact, see [4], that ωb describes a sort of inertia of the system. The "decision probability" is given by the expression: P (A = 1) = lim t→∞ Pt(A = 1) = nb. (2.6) We remark that dependence on the initial belief-state of Alice disappeared. It does not matter whether she was in the belief-state φ0 (she was firmly determined to reply "no") or in the belief- state φ1 (she was firmly determined to reply "yes"). Finally, she set the subjective probability P (A = 1) = nb to reply "yes". Moreover, as we will discuss later, even if initially Alice was in a superposition belief-state (i.e., she started with a belief-state of uncertainty), it would lead to setting of the same probability in favor to reply "yes". This behavior of Alice is natural: she interacts with a homogeneous reservoir, an ensemble of agents, where all agents have the same probability to favor the "yes" answer; Alice simply follows them. In spite of its simplicity, the belief state dynamics of a decision maker interacting with an almost homogeneous information environment can have interesting applications. It describes well not only modern mass-media campaigns on the political arena, but can also be used in economics and finance. 8This approach, which is used everywhere in [4], is the standard way in which the dynamics of any system, micro- or macroscopic, is deduced in classical and in quantum mechanics. What we are doing, in fact, is to adapt the same basic idea to decision making processes. 9The latter assumption is not so unusual. Although the modern information (mental) environment represents huge variety of information flows, an individual is typically coupled to the concrete flow, e.g., she has the custom to see only particular TV-channels and follow only specially selected Internet resources. Even her physical human environment is typically homogeneous. A professor of a university is surrounded by rather homogeneous population with liberal views, including Brexit, Trump or Putin. One of the authors of this paper was visiting USA just before the second Bush-vote. The university environment expressed generally the belief that Bush would not be elected. For such R, the average nb could be estimated as approximately equal to one. 11 For example, in this framework we can model the process of persuasion of consumers. A company wants to persuade consumers to buy some commodity C. So the question A is "to buy or not to buy?" To approach the goal, the company creates an information environment R (almost homogeneous) with the parameter nb ≈ 1. 10 Then our model shows that, for consumers, it is possible to approach the same level of confidence to buy C as in the surrounding information environment R. As was already remarked, the information structure of R need not include directly the question A, i.e., the information campaign need not be straightforwardly oriented to advertising of the commodity C. The R can be concerned with other questions, probably quite different from A. Then the main task of the company driving the persuasion campaign is to establish interaction of consumers with R which can be formally represented by the interaction Hamiltonian considered in this paper. Persuasion of consumers need not be reduced to the concrete commodity C. A group of corporations can perform persuasion regarding a class of commodities, e.g., a new generation of mobile phones, or wind energy, or electric cars. The strategy is the same: creation of an information environment and coupling it with people's opinion in such a way that would generate stabilization of subjective probabilities Pt(A = 1) (in a long run campaign) to the desired value of nb. 3 The stabilization procedure This section is devoted to justification of the model of decision making as decoherence, stabiliza- tion of the belief-state ρ(t) of Alice. In the Heisenberg picture this problem can be transformed into the problem of stabilization of reflection (creation-annihilation) operators a(t), a†(t) and consequently the decision operator na(t) = a†(t)a(t) (the number operator). Of course, such stabilization is possible only for special classes of information (mental) environments and in- teractions between decision maker and surrounding environment. Mathematically the problem is very complicated and we were able to solve it only partially. This section is of the purely mathematical (and quantum physical) nature. If S is a closed quantum system driven by a time-independent, self-adjoint, hamiltonian H, it is natural to suspect that only periodic or quasi-periodic effects can take place, during the time evolution of S. This is because the energy of S is preserved, and this seems to prevent any damping effect in S. For instance, if we work in the Schrodinger representation (SR), the 10Typically this is done through a massive campaign in the mass-media. However, the information content of R is not reduced to direct advertising of C, see considerations below. 12 time evolution Ψ(t) of the wave-function of the system is simply Ψ(t) = e−iHtΨ(0) and, since the operator e−iHt is unitary, we do not expect Ψ(t) to decrease (in some suitable sense) to zero when t diverges. Nevertheless, we will show that a similar decay feature is possible if S is coupled to a reservoir R, but only if R is rather large as compared to S, in particular, if R has an infinite number of degrees of freedom. To see this in details, we start by considering a first system, S, interacting with a second system, S, and we assume for the time being that both S and S are of the same size: to be concrete, this means here that S describes a single particle and, analogously, S describes a second particle. To model dynamics, we introduce the operators a, a† and na = a†a (for the first particle) and the operators b,b† and nb = b†b (for the second particle).11 These operators obey the following canonical commutation relations (CCR): [a, a†] = [b, b†] = I, while all the other commutators are assumed to be zero. A natural choice for a Hamiltonian of the compound system S ∪ S is the following: h = ωana + ωbnb + µ(cid:0)a†b + b†a(cid:1) , where ωa, ωb and µ must be real quantities for h to be self-adjoint. Recall that losing self- adjointness of h would produce a non unitary time evolution, and this is out of the scheme usually considered in ordinary quantum mechanics. The hamiltonian h contains a free part plus an interaction part. The latter is such that, if the eigenvalue of na increases by one unit (during time evolution), then the eigenvalue of nb decreases by one unit, and viceversa. This is because [h, na + nb] = 0, so that na + nb is an integral of motion. Let us now consider dynamics of operators in the Heisenberg representation. The equations of motion for a(t) and b(t), can be easily deduced and turn out to be a(t) = i[h, a(t)] = −iωaa(t) − iµb(t), b(t) = i[h, b(t)] = −iωbb(t) − iµa(t), a(0) = a, b(0) = b. The solution can be written as a(t) = aαa(t) + bαb(t) and b(t) = aβa(t) + bβb(t), where the functions αj(t) and βj(t), j = a, b, are linear combinations of eλ±t, with λ± = −i 2 (ωa + ωb − p(ωa − ωb)2 + 4µ2). Moreover αa(0) = βb(0) = 1 and αb(0) = βa(0) = 0, in order to have 11Cognitive meaning of these operators in the quantum-like model of decision making will be discussed in section 2. 13 a(0) = a and b(0) = b. Hence we see that both a(t) and b(t), and,as a consequence, na(t) = a†(t)a(t) and nb(t) = b†(t)b(t), are linear combinations of oscillating functions, so that no damping is possible within this simple model. Suppose now that the system S is replaced by an (infinitely extended) reservoir R, whose particles are described by an infinite set of CCR-operators b(k), b†(k) and n(k) = b†(k)b(k), k ∈ R. Each k labels one of the elements of the reservoir surrounding S. Hence, for k1 6= k2, we are considering two different elements which may have different characteristics. This is why, in (3.1), we are introducing two k-depending functions, ω(k) and f (k). The hamiltonian of S ∪ R extends h above and is now taken to be H = H0 + λHI, H0 = ωna +ZR where [a, a†] = I, [b(k), b†(q)] = Iδ(k − q), while all the other commutators are zero. It could be useful to notice that here, rather than the CAR used in section 2, we are assuming CCR. We will see later, in section 3.3, that this does not affect our main conclusions. The reason for the different choices is to show that the output of the model does not really depend on the commutation rules adopted. All the constants appearing in (3.1), as well as the regularizing function f (k), are real, so that H = H †. Notice that an integral of motion exists also for S ∪R, na +RR n(k) dk, which extends the one for S ∪ S, na + nb. With this choice of H, the Heisenberg ω(k)n(k) dk, HI =ZR(cid:0)ab†(k) + a†b(k)(cid:1) f (k)dk, equations of motion are (3.1) ( a(t) = i[H, a(t)] = −iωa(t) − iλRR f (k) b(k, t) dk, b(k, t) = i[H, b(k, t)] = −iω(k)b(k, t) − iλf (k) a(t), (3.2) They are supplemented by the initial conditions a(0) = a and b(k, 0) = b(k). In particular, the last equation can be rewritten in an integral form as b(k, t) = b(k)e−iω(k)t − iλf (k)Z t 0 a(t1)e−iω(k)(t−t1) dt1. (3.3) 3.1 An almost homogeneous reservoir In this section we fix f (k) = 1 and ω(k) = ωbk, where ωb ∈ R+. This is a standard choice in quantum optics, [55]. We now insert b(k, t) in (3.3) in the first equation in (3.2), change the order of integration, and use the integral expression for the Dirac deltaRR e−iωbk(t−t1) dk = 2π δ(t− t1), as well as the equalityR t 2g(t) for any test function g(t). Then, we conclude that ωb 0 g(t1)δ(t− t1) dt1 = 1 a(t) = −(cid:18)iω + πλ2 ωb (cid:19) a(t) − iλZR b(k) e−iωbkt dk. (3.4) 14 This equation can be solved, and the solution can be written as a(t) =(cid:18)a − iλZR dk η(k, t)b(k)(cid:19) e−(cid:16)iω+ πλ b (cid:17)t, ω 2 (3.5) ρ(k)(cid:0)eρ(k)t − 1(cid:1) and ρ(k) = i(ω− ωbk) + πλ2 where η(k, t) = 1 . Using complex contour integration it is possible to check, see appendix 2, that [a(t), a†(t)] = I for all t: this means that natural decay of a(t), described in (3.5), is balanced by the reservoir contribution. This feature is crucial since it is a measure of the fact that the time evolution is unitarily implemented in our approach, even if a(t) apparently decays for increasing t. ωb Let us now consider a state over S ∪R, h XS ⊗ XRi = hϕna, XSϕnai h XRiR, in which ϕna is an eigenstate of the number operator na and < >R is a state of the reservoir, which is assumed to satisfy, among other properties (see [55, 4]), (cid:10) b†(k)b(q)(cid:11)R = nb(k)δ(k − q). (3.6) This is a standard choice, see for instance [55], which extends the choice we made for S. Here XS ⊗ XR is the tensor product of an operator of the system, XS, and an operator of the reservoir, XR. Then, if for simplicity we take the function nb(k) to be constant in k we get, calling na(t) :=< na(t) >=< a†(t)a(t) >, na(t) = na e− 2λ ωb 2 π t + nb(cid:18)1 − e 2λ 2 π ωb t(cid:19) , (3.7) which goes to nb as t → ∞. Hence, if 0 ≤ nb < na, the value of na(t) decreases with time. If, on the other hand, nb > na, then the value of na(t) increases for large t. This is the exponential rule which, as discussed before, cannot be deduced if R has not an infinite number of degrees of freedom. Notice that, in particular, if the reservoir is originally empty, nb = 0, then na(t) = na e− 2λ t decreases exponentially to zero: the system becomes empty. On the ω π 2 b other hand, since na +RR n(k) dk is a constant of motion, the reservoir starts to be filled up. Remark 1. The continuous reservoir considered here (k ∈ R) could be replaced by a discrete one, describing again an infinite number of particles, but labeled by a discrete index. In this case, to obtain a Dirac delta distribution, which is the crucial ingredient in the derivation above, we have to replace the integralRR e−ik(t−t1) dk = 2πδ(t− t1) with the Poisson summation formula, which we write here as Pn∈Z einxc = 2π In the above model we see that bath is essentially characterized by three functions: f (k), ω(k) and nb(k). In particular, in what we have done so far, we have taken f (k) = 1, ω(k) = ωb k c Pn∈Z δ(cid:0)x − n 2π c (cid:1), for all non zero c ∈ R. 15 and nb(k) = nb. This choice can be interpreted as follows: the bath is almost homogeneous, meaning that the functions f (k) and nb(k), which in principle could depend on k (remember that each k labels an element of the bath), are constant in k: different members of the bath all share the same values of f (k) and of nb(k). However, these members are not all completely identical, since ω(k) depends on k, as we have already observed in section 2. 3.2 Changing reservoir To keep the situation under control as much as possible, we now discuss what happens if we keep f (k) and ω(k) fixed as before, but now allow for a different dependence of nb(k) on k. In other words, we do not assume that nb(k) is constant in k, while we still take f (k) = 1 and ω(k) = ωb k. With this choice, a(t) is given again as in (3.5), and we conclude, in the same way, that [a(t), a†(t)] = I. The function na(t) is the one in (4.5) but, this time, we cannot use (4.4) since nb(k) is no longer a constant function. The computation of RR nb(k)η(k, t)2dk is quite similar to that of RR η(k, t)2dk, with the obvious difference due to the presence of nb(k) inside the integral. What is essential for us is that the integral is real-valued for all real choices of nb(k). To use again the complex integration techniques, it is useful to assume that nb(k) is analytic in k, and that nb(k) does not diverge when k diverges. For concreteness, we take (3.8) nb(k) = nb k2 + α2 , for some positive nb and for some α > 0. Notice that we could safely take α < 0, too; our choice is fixed only for demonstration. On the other hand, the condition nb > 0 is essential because it follows from the origin of nb(k), which came from a positive operator. The computation of RR nb(k)η(k, t)2dk is performed again with the techniques discussed in details in appendix 2, with the only difference that we have two singularities of the integrating functions both in the upper and in the lower complex semi-planes, so the result of the integral is the sum of two residues. The computation can be performed thoroughly, and an analytic form of na(t) could be given for all t. However, what is more relevant for us is large time limit of na(t), which turns out to be na(∞) := lim t,∞ na(t) = . (3.9) Notice that this result makes sense only if α and λ are not zero. This suggests that the complex pole in nb(k), and the interaction between Alice and her bath, are really essential to get some nb(αω2 b + πλ2) +(cid:16)α + πλ2 ω2 b (cid:17)2(cid:19) αλ2ω2 b(cid:18) ω2 ω2 b 16 stabilization. We observe also that the asymptotic value of the decision function na(t) strongly depends on the various parameters of the model, which could be adjusted to fit experimental data, if needed. Notice that the nb(k) in (3.8) describes a bath which is far from being uniform. Here, in particular, nb(k) → 0 for k → ∞. Hence not all the parts of the bath interact with Alice in the same way: serious differences arise. Remark 2. In this section we have considered a bath which the amplitude nb(k) goes to zero as k−2, see (3.8). In fact, we could think of a function nb(k) decreasing like k−1, but this is not compatible with the reality of the function nb(k), at least if we still want to use complex integration techniques to compute na(t). The reason is that, to avoid singularities in the real axis, we are forced to choose nb(k) = µ k−iα, for some complex µ and for some α > 0. This ωb (cid:17)(cid:17), for some produces na(∞) = γωb, at least if µ is chosen as follows: µ = γ(cid:16)ω − i(cid:16)αωb + πλ2 real γ. So we see that, with this peculiar choice of nb(k), we still get a real na(∞), but what we cannot exclude that for finite time na(t) can be complex. Remark 3. Both (3.7) and (3.9) suggest that the asymptotic limit of na(t) does not really depend on whether, at t = 0, the system is in a pure state or in a combination of states. This was already anticipated in section 2. We are using the Heisenberg representation. This means that the state does not change, while the observables depend on time. 3.2.1 A remark on f (k) and ω(k) So far we have fixed f (k) = 1 and ω(k) = ωb k. This is not the only natural choice, in particular if we want to stress the difference between different parts of the reservoir. What is technically useful for us, is that Dirac's delta function appears when deducing the differential equation for a(t), see appendix 2. In fact, this is also possible under other conditions. Suppose that the two functions f (k) and ω(k) are such that: ω(k) → ±∞ for k → ±∞, and dω(k) dk = 1 β f 2(k), for some real β 6= 0. These assumptions are satisfied if f (k) = 1 and ω(k) = ωbk, but also in many other situations. Hence ZR f 2(k)e−iω(k)(t−t1) dk = βZR e−iω(k)(t−t1) dω(k) = 2πβδ(t − t1), 17 and we can significantly simplify the equation for a(t), which now becomes a(t) = −(cid:0)iω + βπλ2(cid:1) a(t) − iλZR f (k)b(k) e−iω(k)t dk. (3.10) This is the equation which replaces (3.4) in this new situation. The solution can be found as before, and we get a(t) =(cid:18)a − iλZR dk η(k, t)b(k)f (k)(cid:19) e−(iω+βπλ2)t, (3.11) ρ(k)(cid:0)eρ(k)t − 1(cid:1) and ρ(k) = i(ω − ω(k)) + βπλ2. At a first sight, the situation where η(k, t) = 1 is not particularly different from those discussed earlier and in appendix 2. However, a serious technical problem now arises: to find [a(t), a†(t)] and na(t), we need to compute integrals and the poles of the integrating functions are strongly dependent on the analytic expression of ω(k): in particular, if ω(k) is not linear, the computations become rather hard! 3.3 What if we use CAR? In the Hamiltonian (3.1) the operators a and b(k) are assumed to satisfy CCR. However, from the point of view of a DM procedure, it might be more interesting to consider the case in which Alice's decisions are described by CAR-operators. This means that, first of all {a, a†} = a a† + a†a = I, with {a, a} = 0. It is then natural to consider a CAR-bath as well, i.e. to assume that the operator b(k) satisfies the following rules: {b(k), b†(q)} = δ(k − q) I, {b(k), b(q)} = 0, {a♯, b♯(k)} = 0, x♯ being either x or x†. Assuming the same Hamiltonian (3.1), differential equations of motion for the annihilation operators a(t) and b(k, t) can be deduced. They turn out to be the same as in (3.2), except that λ must be replaced by −λ. In other words, we get ( a(t) = i[H, a(t)] = −iωa(t) + iλRR f (k) b(k, t) dk, b(k, t) = i[H, b(k, t)] = −iω(k)b(k, t) + iλf (k) a(t). (3.12) As in the CCR-case, we assume that in (3.6) the function nb(k) is constant, nb(k) ≡ nb ≥ 0. The CAR-setting implies that this constant cannot exceed one, 0 ≤ nb ≤ 1 (this is the average of the CAR-number operator). As we can see from (3.7), for the almost homogeneous reservoir, that na(t) depends on λ2 and not on λ. The same is true also for the choice (3.8) of the reservoir. This is evident from 18 (3.9) for large values of t, but can also be checked for any finite t. In other words, even with this different choice of nb(k), we observe that na(t) depends on λ2 and not on λ. The conclusion, therefore, is clear: the CCR-CAR nature of our model does not affect the main results we have deduced. One choice or the other should be related to the nature of the decision we want to model: in case of a binary question (yes or not), the natural settings is probably the CAR- one. But if we assume that several possible answers (normally, infinite!) are possible, then we should use a CCR-version of the model. In this case, the existence of a quadratic integral of motion can limit the number of possible answers to the original questions, giving rise to a sort of finite-dimensional Hilbert space, similarly to what is discussed in details in [24]. 4 Concluding discussion In modeling of the process of decision making, the main output of this paper is the description of properties of an information reservoir (environment) R leading to asymptotic stabilization. In other words, this is a description of a context surrounding an agent, Alice, which guaranties that she would make some decision, i.e., her mind would not fluctuate for ever between a variety of alternative answers to a question (solutions of a problem) A. The presented results on asymptotic stabilization can be used to support mathematically the model of decision making as decoherence - in its quantum field version. Here the theory of open quantum systems is applied straightforwardly in the Hamiltonian framework. Decision making dynamics is given by the Heisenberg equations for operators describing reflections of an agent, Alice, and the operators representing the information reservoir. The most interesting for decision making are constraints imposed on information reservoirs, environments, guaran- teeing asymptotic stabilization. Our present study is only a first step in this direction; further studies of sufficient and necessary conditions of asymptotic stabilization are needed.12 We hope that this paper will attract attention of experts in mathematical physics (especially working on mathematical problems of quantum field theory) to the problem of asymptotic stabiliza- tion. Novel applications to social science, microeconomics, and finance, generate new laws for functions f (k), nb(k), ω(k) which have not been considered in physical applications. 12 We remark that applications of quantum field formalism to modeling of decision making generate new dynamical models based on algebras of qubit creation-annihilation operators [57]. These operators satisfy the so called qubit commutation relations which combine both CAR and CCR-features. To the best of our knowledge, there are no results concerning asymptotic stabilization for such dynamical processes and this is an interesting topic for research. 19 To extend the readership of this paper, the calculations in section 3 were done at the "phys- ical level of rigor". A complete mathematical treatment of the problem should involve consider- ation of operator-valued distributions (generalized functions). Such treatment definitely must be performed at some stage of future development of quantum field inspired models of decision making. In general, the great success of quantum field theory of physics rises the expectations that this formalism will fruitfully contribute to modeling of decision making and cognitive processes. Appendix 1: Second quantization: CAR We briefly review some basic facts on the so–called canonical anti-commutation relations (CAR), which were originally introduced in connection with what in quantum physics is called sec- ond quantization for identical particles with half-integer spin. We say that a set of operators {aℓ, a† ℓ, ℓ = 1, 2, . . . , L} satisfy the CAR if the conditions {aℓ, a† n} = δℓnI, {aℓ, an} = {a† ℓ, a† n} = 0 (4.1) hold true for all ℓ, n = 1, 2, . . . , L. Here, I is the identity operator and {x, y} := xy + yx is the anticommutator of x and y. These operators are those which are used to describe L different modes of fermions. From these operators we can construct nℓ = a† ℓ=1 nℓ, which are both self–adjoint. In particular, nℓ is the number operator for the ℓ–th mode, while N is the number operator of S. Compared with bosonic operators, the operators introduced here satisfy a very important feature: if we try to square them (or to rise to higher powers), we simply get zero: for instance, from (4.1), we have a2 ℓ = 0. This is related to the fact that fermions satisfy the Fermi exclusion principle. ℓaℓ and N =PL The Hilbert space of our system is constructed as follows: we introduce the vacuum of the theory, that is a vector ϕ0 which is annihilated by all the operators aℓ: aℓϕ0 = 0 for all ℓ = 1, 2, . . . , L. Then we act on ϕ0 with the operators a† ℓ (but not higher powers, since these powers are simply zero!): ϕn1,n2,...,nL := (a† 1)n1(a† 2)n2 · · · (a† L)nLϕ0, (4.2) nℓ = 0, 1 for all ℓ. These vectors give an orthonormal set and are eigenstates of both nℓ and N: ℓ=1 nℓ. Moreover, nℓϕn1,n2,...,nL = nℓϕn1,n2,...,nL and Nϕn1,n2,...,nL = Nϕn1,n2,...,nL, where N = PL using the CAR, we deduce that nℓ (aℓϕn1,n2,...,nL) = (nℓ − 1)(aℓϕn1,n2,...,nL) 20 and nℓ(cid:16)a† ℓϕn1,n2,...,nL(cid:17) = (nℓ + 1)(a† l ϕn1,n2,...,nL), for all ℓ. Then aℓ and a† ℓ are called the annihilation and the creation operators. In fact, they transform the vector ϕn1,n2,...,nℓ,...,nL into a different vector, proportional to ϕn1,n2,...,nℓ±1,...,nL, annihilating or creating one particle in the ℓ-th mode. However, in some sense, a† ℓ is also an annihilation operator since, acting on a state with nℓ = 1, we destroy that state. The Hilbert space H is obtained by taking the linear span of all these vectors. Of course, H has a finite dimension. In particular, for just one mode of fermions, dim(H) = 2. This also implies that, contrarily to what happens for bosons, the fermionic operators are bounded. The vector ϕn1,n2,...,nL in (4.2) defines a vector (or number) state over the algebra A as ωn1,n2,...,nL(X) = hϕn1,n2,...,nL, Xϕn1,n2,...,nLi, (4.3) where h , i is the scalar product in H. Appendix 2: Explicit computations In this appendix we give some details of our computations just sketched in Section 3.1, useful for those who are not familiar with this kind of interacting open systems. Our starting point is b(k, t) in (3.3), with f (k) = 1 and ω(k) = ωb k. Hence b(k, t) = b(k)e−iωbkt − iλZ t 0 a(t1)e−iωbk(t−t1) dt1. When we replace this expression in the equation for a(t), a(t) = −iωa(t) − iλRR b(k, t) dk, we have to compute, in particular, the double integral ZR(cid:18)Z t 0 a(t1)e−iωbk(t−t1) dt1(cid:19) dk =Z t 0 a(t1)(cid:18)ZR a(t1)δ(t − t1)dt1 = π ωb e−iωbk(t−t1) dk(cid:19) dt1 = a(t), = 2π ωb Z t 0 and (3.4) now follows. An easy way to solve (3.4) consists in using the change of variable a(t) = A(t) expn−(cid:16)iω + πλ2 ωb (cid:17) to since, in terms of A(t), (3.4) can be rewritten as A(t) = −iλZR b(k) e−iωbkt dk, ⇒ A(t) = A(0) − iλZ t 0 (cid:18)ZR b(k) e−iωbkt1 dk(cid:19) dt1, 21 which returns, after few simple computation, the solution in (3.5). To check now that the CCR are preserved during time evolution, i.e. that [a(t), a†(t)] = I for all t, we observe that the operators at the right hand side of (3.5) are the initial (t = 0) operators, so that they satisfy the commutation rules [a, a†] = I, [b(k), b†(q)] = δ(k − q), while [a, b(k)] = [a†, b(k)] = ... = 0. Therefore [a(t), a†(t)] = e− 2πλ ωb 2 t(cid:18)1 + λ2ZR η(k, t)2dk(cid:19) I. In our case, after some simple algebra, we deduce that ZR η(k, t)2dk =ZR ω2 ωb ± i πλ2 where k± = ω we conclude that ω2 b dk b (k − k+)(k − k−)(cid:20)(cid:18)e 2 2πλ ω b 2 t + 1(cid:19) − e(−iω+ πλ ω b )teiωbkt − e(iω+ πλ ω b 2 )te−iωbkt(cid:21) , . Using complex integration we can compute each contribution here and ZR η(k, t)2dk = 2 2πλ ωb 1 λ2 (cid:18)e t − 1(cid:19) , (4.4) so that our claim easily follows. To prove now (3.7), we first observe that the mean values of contributions like a†b(k) and ab†(k) on our states are zero. Therefore, using (3.5), we get na(t) = e− 2πλ ω b 2 t(cid:18)(cid:10)ϕna, a†aϕna(cid:11) h IRiR + λ2 hϕna, IϕnaiZRZR 2 η(k, t)η(q, t) (cid:10) b†(k)b(q)(cid:11)R dk dq(cid:19) = Then, if nb(k) = nb, we can use the result in (4.4) and recover the result in (3.7). = e− 2πλ ω b t(cid:18)na + λ2ZR nb(k)η(k, t)2dk(cid:19) . (4.5) Acknowledgments. This paper is based on the discussions between the coauthors during the visit of two of them (IB and KHR) to the university of Palermo in April 2017. IB was supported by Marie Curie Fellowship at City University of London, H2020-MSA-IF-2015, grant N 696331; AKH was supported by the EU-project Quantum Information Access and Retrieval Theory (QUARTZ), Grant No. 721321. FB thanks the Gruppo Nazionale di Fisica Matematica of I.N.d.A.M., and by the University of Palermo, for support. F.B. also thanks the Distinguished Visitor Program of the Faculty of Science of the University of Cape Town, 2017. References [1] A. Khrennikov, Information dynamics in cognitive, psychological, social, and anomalous phenomena. Ser.: Fundamental Theories of Physics, Kluwer, Dordreht, 2004. 22 [2] A. Khrennikov, Ubiquitous quantum structure: from psychology to finance. Springer, Berlin-Heidelberg-New York, 2010. [3] Busemeyer J R, Bruza P D. 2012 Quantum models of cognition and decision. Cambridge: Cambridge University Press. [4] Bagarello F. 2012 Quantum dynamics for classical systems: with applications of the number operator. New York: J. Wiley. [5] Haven E, Khrennikov A. 2013 Quantum social science. Cambridge: Cambridge University Press. [6] Asano M, Khrennikov A, Ohya M, Tanaka Y, Yamato I. 2015 Quantum adaptivity in biology: from genetics to cognition. Heidelberg-Berlin-New York: Springer. [7] Khrennikov, A., 2004. On quantum-like probabilistic structure of mental information. Open Systems and Information Dynamics 11 (3), 267-275. [8] Khrennikov, A., 2006. Quantum-like brain: Interference of minds. BioSystems 84, 225-241. [9] Busemeyer J R, Wang Z, Townsend J T. 2006 Quantum dynamics of human decision making. J. Math. Psychol. 50,220-241 [10] Pothos E M, Busemeyer J R. 2009 A quantum probability explanation for violation of rational decision theory. P. Roy. Soc. Lond. B Bio. 276, 2171-2178 [11] Busemeyer J R , Pothos E M, Franco R, Trueblood J. 2011 A quantum theoretical expla- nation for probability judgment errors. Psychol. Rev. 118, 193-218 [12] Dzhafarov, E.N. & Kujala, J.V. 2012 Selectivity in probabilistic causality: Where psychol- ogy runs into quantum physics. Journal of Mathematical Psychology 56, 54-63. [13] Asano M, Ohya M, Tanaka Y, Basieva I, Khrennikov A. 2012 Quantum-like dynamics of decision-making. Physica A 391, 2083-2099 [14] Hawkins R J and Frieden B R. 2012 Asymmetric Information and Quantization in Financial Economics. Int. J. Math. and Math. Sc. 32012, Article ID 470293. [15] Pothos E M , Busemeyer J R.2013 Can quantum probability provide a new direction for cognitive modeling? Behav. Brain Sci. 36, 255–327 23 [16] Asano M, Basieva I, Khrennikov A, Ohya M, Tanaka Y, Yamato I 2013. A model of epigenetic evolution based on theory of open quantum systems. Systems and Synthetic Biology 7, 161-173 [17] Dzhafarov E N and Kujala J V 2014 On selective influences, marginal selectivity, and Bell/CHSH inequalities. Topics in Cognitive Science 6, 121-128 [18] Kvam P D, Busemeyer J R, Lambert-Mogiliansky A. 2014 An empirical test of type- indeterminacy in the Prisoners Dilemma. Lecture Notes in Computer Science 8369,213- 224 [19] Khrennikov A, Basieva I. 2014 Possibility to agree on disagree from quantum information and decision making. J. Math. Psychol. 62–63, 1–15 [20] Plotnitsky A. 2014 Are quantum-mechanical-like models possible, or necessary, outside quantum physics? Phys. Scripta T163, 014011 [21] A. Khrennikov, Basieva, I. (2014). Quantum Model for Psychological Measurements : From the Projection Postulate to Interference of Mental Observables Represented As Positive Operator Valued Measures. NeuroQuantology. 12. 324-336 [22] A. Khrennikov, Basieva, I., Dzhafarov, E.N., Busemeyer, J.R. 2014. Quantum Models for Psychological Measurements : An Unsolved Problem. PLoS ONE. 9. Article ID: e110909. [23] J. R. Busemeyer, Z. Wang, A. Khrennikov, I. Basieva, I. 2014. Applying quantum principles to psychology. Physica Scripta. T163. 014007. [24] Bagarello F., Oliveri F., An operator-like description of love affairs, SIAM Jour. Appl. Math., 70, No. 8, pp. 32353251, (2010) [25] Bagarello F., Gargano F., Modeling interactions between political parties and electors, Phys. A, dx.doi.org/10.1016/j.physa.2017.04.035 [26] Bagarello F. 2015 An operator view on alliances in politics. SIAM J. Appl.. Math. 75(2), 564-584 [27] Bagarello F, Haven E. 2016 First results on applying a non-linear effect formalism to alliances between political parties and buy and sell dynamics. Physica A 444, 403-414 24 [28] Bagarello F. 2015 A quantum-like view to a generalized two players game. Int. J. Theor. Phys. 54 (10), 3612-3627 [29] Denolf J and Lambert-Mogiliansky A. 2016. Bohr complementarity in memory retrieval. J. Math. Psych. 73, 28-36 [30] E. Haven and A. Khrennikov, Statistical and subjective interpretations of probability in quantum-like models of cognition and decision making. J. Math. Psych. 74, 8291 (2016). [31] R. J. Hawkins and B. R. Frieden, 2017 Quantization in financial economics: An information-theoretic approach. The Palgrave handbook of quantum models in social sci- ence. Applications and Grand Challenges. Haven E and Khrennikov A (eds). Palgrave Macmillan UK, pp. 19-38. [32] Lambert-Mogiliansky A. 2017. Quantum-like type indeterminacy: A constructive approach to preferences a la Kahneman and Tversky. Haven E and Khrennikov A (eds). Palgrave Maxmilan UK, pp. 229-250. [33] Danilov V I, Lambert-Mogiliansky A and Vassili Vergopoulos V. 2016 Dynamic consistency of expected utility under non-classical(quantum) uncertainty. https://halshs.archives- ouvertes.fr/halshs-01324046/ [34] A. Khrennikov, Classical and quantum mechanics on information spaces with applications to cognitive, psychological, social and anomalous phenomena. Found. of Physics, 29, N. 7, 1065-1098 (1999). [35] M. Asano, I. Basieva, A. Khrennikov, M. Ohya, Y. Tanaka, I. Yamato, Quantum infor- mation biology: from information interpretation of quantum mechanics to applications in molecular biology and cognitive psychology. Found. Phys. 45, N 10, 1362-1378 (2015). [36] D' Ariano, G. M.: Physics as information processing. In: Jaeger, G. et al. (eds.), Advances in Quantum Theory, AIP, Melville, NY, 1327, pp. 7–16 (2011) [37] Plotnitsky, A.: Niels Bohr and Complementarity: An Introduction. Springer, Berlin and New York (2012) [38] Penrose, R.: The Emperor's new mind. Oxford Univ. Press, New-York (1989) [39] Hameroff, S.: Quantum coherence in microtubules. A neural basis for emergent conscious- ness? J. of Cons. Stud. 1, 91-118 (1994) 25 [40] Tegmark, M. (2000). Importance of quantum decoherence in brain processes. Phys. Rev. E 61 (4): 41944206 [41] de Barros A J , Suppes P. 2009 Quantum mechanics, interference, and the brain. J. Math. Psychol. 53, 306-313. [42] A. Khrennikov, Quantum-like model of processing of information in the brain based on classical electromagnetic field. Biosystems 105(3), 250-262 (2011). [43] de Barros, A. J. Quantum-like model of behavioral response computation using neural oscillators. Biosystems 110, 171-182 (2012) [44] Takahashi T. Toward a physical theory of quantum cognition. Topics in Cognitive Science 6, 104107 (2014). [45] Melkikh, A.V., 2013. Biological complexity, quantum coherent states and the problem of efficient transmission of information inside a cell. BioSystems. 111, 190-198. [46] Melkikh, A.V., 2014. Congenital programs of the behavior and nontrivial quantum effects in the neurons work. BioSystems. 119, 10-19. [47] Busemeyer, J. R., Fakhari, P., Kvam, P. 2017. Neural implementation of operations used in quantum cognition. Progress in Biophysics and Molecular Biology, in press. [48] von Neumann J., Mathematical Foundations of Quantum Mechanics, Princeton Univ. Press, Princenton, 1955. [49] Basieva I and Khrennikov A. 2017 Decision-making and cognition modeling from the theory of mental instruments. The Palgrave Handbook of Quantum Models in Social Science. Applications and Grand Challenges. Haven E and Khrennikov A (eds). Palgrave Maxmilan UK, pp. 75-93. [50] Khrennikov A. 2017 Why quantum? The Palgrave Handbook of Quantum Models in Social Science. Applications and Grand Challenges. Haven E and Khrennikov A (eds). Palgrave Maxmilan UK, pp. 321-334. [51] Asano M, Ohya M, Tanaka Y, Basieva I, Khrennikov A. 2011 Quantum-like model of brain's functioning: decision making from decoherence. J. Theor. Biol. 281, 56-64 26 [52] M. Asano, M. Ohya, A. Khrennikov, Quantum-Like Model for Decision Making Process in Two Players Game - A Non-Kolmogorovian Model. Found. Phys., 41, 538-548 (2011). [53] Skomski R, Kashyap A and Sellmyer D J 2012 A quantum-mechanical relaxation model. J. Appl. Phys.111, 07D507 [54] Bagarello F, Haven E., Khrennikov A., A model of adaptive decision making from repre- sentation of information environment by quantum fields. Phil. Trans. A, to be published in 2017. [55] S.M. Barnett, P.M. Radmore, Methods in theoretical quantum optics, Clarendon Press, Oxford, 1997 [56] A. Khrennikov, "Social laser:" action amplification by stimulated emission of social energy. Phil. Trans. Royal Soc. 374, N 2054, 20150094 (2016). [57] P. Khrennikova, Modeling behavior of decision makers with the aid of algebra of qubit creation-annihilation operators J. Math. Psych. published on line (2016) 27
0902.2020
3
0902
2011-11-01T01:49:42
Feature selection in simple neurons: how coding depends on spiking dynamics
[ "q-bio.NC", "q-bio.QM" ]
The relationship between a neuron's complex inputs and its spiking output defines the neuron's coding strategy. This is frequently and effectively modeled phenomenologically by one or more linear filters that extract the components of the stimulus that are relevant for triggering spikes, and a nonlinear function that relates stimulus to firing probability. In many sensory systems, these two components of the coding strategy are found to adapt to changes in the statistics of the inputs, in such a way as to improve information transmission. Here, we show for two simple neuron models how feature selectivity as captured by the spike-triggered average depends both on the parameters of the model and on the statistical characteristics of the input.
q-bio.NC
q-bio
Feature selection in simple neurons: how coding depends on spiking dynamics Michael Famulare University of Washington Department of Physics Box 351560 Seattle, WA 98195-1560 [email protected] Adrienne Fairhall University of Washington Department of Physiology and Biophysics HSB G424, Box 357290 Seattle, WA 98195-7290 October 29, 2018 Abstract The relationship between a neuron's complex inputs and its spiking output defines the neuron's coding strategy. This is frequently and ef- fectively modeled phenomenologically by one or more linear filters that extract the components of the stimulus that are relevant for triggering spikes, and a nonlinear function that relates stimulus to firing proba- bility. In many sensory systems, these two components of the coding strategy are found to adapt to changes in the statistics of the inputs, in such a way as to improve information transmission. Here, we show for two simple neuron models how feature selectivity as captured by the spike-triggered average depends both on the parameters of the model and on the statistical characteristics of the input. Neuronal dynamics are characterized by nonlinearities that lead to large, approximately stereotyped voltage excursions, or spikes, that are the basis for interneuronal signaling. Capturing the relationship between inputs and the resulting pattern of spike outputs from a given neuron in the form of a reduced functional model is a focus of sensory neuroscience. In the sense 1 that such a model provides a general mapping from input to output, it can be thought of as the neuron's "coding strategy". Reverse correlation methods (Bryant and Segundo, 1976; de Boer and Kuyper, 1968; Sakai, 1992; Hunter and Korenberg, 1986) provide a means to sam- ple the statistical characteristics of stimuli that tend to trigger spikes; in the simplest case, the mean, or spike-triggered average stimulus (STA), is the optimal linear kernel for predicting the firing rate from the stimulus (Rieke et al., 1996). Using reverse correlation, one may obtain an approx- imate functional model for the neuronal input/output transformation in terms of the input features that drive the system (Meister and Berry II, 1999; de Ruyter van Steveninck and Bialek, 1988; Simoncelli et al., 2004). These methods may be applied not only to determine how neural systems are driven by external stimuli, but to extract a model for how specific pat- terns of synaptic current inputs drive single neurons. This allows one to de- termine the role that a single neuron with a characteristic complement of ion channels plays in a circuit: the integration of inputs over a certain timescale (Slee et al., 2005; Svirskis et al., 2003; Prescott et al., 2006), the detection of sudden change or highly synchronous events (Abeles, 1982; Slee et al., 2005; Svirskis et al., 2004), or the selection of certain frequency components in the input (Izhikevich, 2001; Prescott et al., 2006). Here, we will derive explicit expressions for the outcome of such a statis- tical analysis applied to two simple neuron models. We have two goals. The first is to develop a general framework for understanding how the details of neuronal dynamics establish or influence the features in the input that trigger spikes. Second, neuronal systems show adaptation to statistics, in the sense that the neuron's coding strategy often changes when driven by stimuli with different statistical properties. In the case of single neurons, such effects can modulate or gate the effective computation of the neuron according to the statistical properties of the signal or the background inputs (Hasenstaub et al., 2007; Destexhe and Pare, 1999; Fellous et al., 2003). To identify the rules governing this process, one would like to know to what ex- tent the observed changes may result from time-independent neuronal non- linearities and to what extent they must be due to changes in underlying neuronal parameters. To study this, we will compute how the experimentally obtained features of two fixed models depend on the statistical properties of the stimulus, focusing on the variance of a white noise input. The key points of this paper are: • The relevant linear filter corresponding to a nonlinear spiking neuron model is determined by the nonlinear dynamics linearized in a manner 2 consistent with the typical operating regime of the system, which is determined both by its dynamics and by the stimulus conditions. To characterize this regime, we compute the voltage probability distribu- tions from the Fokker-Planck interpretation of the models. • We then use a novel application of the technique of stochastic lin- earization to map the nonlinear models onto a set of linear models. By studying both the mapping, determined by an optimization func- tion relating the linear and nonlinear models, and the related STA predictions for the equivalent linear system, we can delineate the roles of different nonlinearities on spike encoding. • The form of the STA is influenced both by the subthreshold (non)linear dynamics and the spike afterhyperpolarization. • Models with similar phase space topology can have STAs whose form is controlled by different mechanisms. A rapid-onset exponential integrate- and-fire model (EIF) has no significant subthreshold nonlinearity, and so its STA is almost completely determined by the probability current due to spiking. In contrast, the quadratic integrate-and-fire model (QIF), while superficially similar to the EIF, has an STA whose form is dominated by the sampling of the subthreshold nonlinearity, with spiking effects playing a secondary role. 1 Models and numerical methods Change in the effective feature selectivity with driving variance has been studied for the case of the leaky integrate-and-fire (LIF) model (Paninski et al., 2003; Paninski, 2006b; Yu and Lee, 2003). In the LIF model, the dynamics are linear until the voltage reaches an imposed threshold after which the voltage is immediately reset below threshold. Thus, the LIF contains no intrinsic excitability, and further, does not allow for the possibility that the system can cross threshold multiple times before spiking due to noisy inputs. This discontinous behavior with respect to spike initiation is not found in biological neurons. Two simple models with more realistic spike initiation are the quadratic and exponential integrate-and-fire models (QIF and EIF, respectively) (Ermentrout and Kopell, 1986; Fourcaud-Trocme et al., 2003). Both models are similar in spirit to the LIF insofar as they replace the afterhyperpolarization mechanism with a discontinous jump, or after-spike reset, but the point of reset in these 3 models occurs at the peak of the spike instead of at the threshold voltage. This mitigates the effects of the pathological behavior in response to noise that the discontinuity creates (Paninski, 2006b) by moving it away from the interesting region of spike initiation. The models are described by an equation of the form: τm v = −v + f (v) + (Vr − Vs) δ(v − Vs) + s(t), (1) where v denotes the membrane voltage, τm is the membrane time constant, Vs is the voltage that defines the spike height and Vr is the post-spike reset voltage. The input current, s(t), is a zero-mean gaussian white noise (GWN) process with correlation function hs(t)s(0)i = σ2τmδ(t). The delta-function is shorthand for the act of resetting the voltage to Vr after it reaches Vs. All of the spike-generating and nonlinear subthreshold dynamics are encoded in f (v). For the two models studied here, we have: f (v) =(αv2 g exph v−V∗ g i for the QIF, for the EIF. (2) We study parameters such that the resting potential is zero and the un- stable fixed point is at α−1 for both models. This requires us to choose V∗ = α−1 (1 + gα ln(gα)) and g ≪ V∗. Somewhat paradoxically, despite the higher order nonlinearity, the choice of small g causes the exponential non- linearity to turn on over a much tighter range in voltage than the quadratic nonlinearity of the QIF. We will see that this leads to more linear behavior of the EIF model below threshold. Thus the two models behave noticeably differently below threshold while still having the same after-spike dynamics. 1.1 Reverse correlation analysis Reverse correlation is used to determine characteristics of the stimulus that are correlated with neuronal response. From a long, random stimulus pre- sentation s(t) and the resulting spike response times ti, one collects the set of N current traces that led to a spike, s(τ − ti), over an interval of time τ = [0, −T ] prior to the spike where T is chosen appropriately to capture all the stimulus history that is relevant to triggering the spike. The spike- triggered average or STA, ¯s(τ ), is found by averaging these samples over i: N ¯s(τ ) = 1 N s(τ − ti). (3) Xi=1 4 1.2 Defining spike times The results of reverse correlation analysis can depend on how the spike time is defined. Here, we will look at the STAs with a temporal resolution that is short compared to the average spike width. Different choices of the voltage threshold used to define spikes will accordingly lead to STAs that differ from each other due to temporal jittering of the ensemble of spike- triggered trajectories. Because we want to understand how the spiking of the model determines the feature selected from the stimulus ensemble, we are interested in choosing a threshold that yields an STA that best captures the role of the stimulus on the approach to the spike but is not sensitive to stimulus-driven variations in the spike itself. For both models considered here, this is achieved by selecting the unstable fixed point that separates the subthreshold region from the spiking region in the absence of noise. Since the location of the unstable fixed point is a function of the mean input current and the quadratic form of the nonlinearity, we will call it the dynamical threshold in accordance with previous work (Izhikevich, 2000; Hong et al., 2007). For the zero-mean inputs considered here, the dynamical threshold is Vth = α−1. Thus, we define spike times as the time of the last upward crossing of the dynamical threshold preceding an after-spike reset. 1.3 Model simulation In discrete-time with time step h, the nonlinear models in equation 1 were realized as: vn+1 = vn + h τm (−vn + f (vn)) + σr h τm if vn+1 ≥ Vs, then vn+1 → Vreset . ξn , (4) (5) where the ξn are drawn from a gaussian distribution with zero mean and unit variance. For all figures in this paper, simulations were run with a time step of h = τm/200 until 2 × 105 spikes were accumulated. The noise was generated with randn in Matlab R2007b. Parameters used in simulation: α = 1, τm = 1, Vs = 25, Vr = −0.2, g = 1 10 , and V∗ = 0.77. 2 Numerical results We computed the STA numerically for a range of values of the stimulus standard deviation σ for both models. Results are shown in figure 1. The STA at all values of σ has two components: an extended feature and a sharp 5 upward step at the time of the spike. For the feature, two different types of behavior appear. For large σ, the STAs are approximately decaying expo- nentials for which, as the standard deviation increases, the decay timescale decreases and the amplitude increases. Thus, at larger σ, the QIF and EIF models perform approximately linear leaky integration, where the effective leakiness depends on the standard deviation. For very small σ, the STAs are non-monotonic, with the peak amplitude occurring well before the spike time. 3 Approximate STA for finite standard deviations To best understand how details of the models influence the STAs, we would like to be able to calculate the STAs analytically. In the zero standard deviation limit, the STA can be analytically calculated for the QIF via a large deviations principle and path integral methods (Paninski, 2006a; Badel et al., 2008a; Wilson and Steyn-Ross, 2008), but that type of analysis does not extend to finite σ. However, path integral methods can be applied for arbitrary σ to perfectly linear models with no reset. As noted previously (Hong et al., 2007), the observation that the STA is an exponential implies that the subthreshold dynamics of the model are effectively linear. Since the STAs of the nonlinear models are roughly exponential for larger standard deviations, we should be able to introduce a linear approximation to the nonlinear models that captures the qualitative behavior of the STA and helps explain in detail how the STA arises from the form of the nonlinearity. The main idea is as follows. While it is impossible to derive complete, time-dependent statistical distributions for these models, we can get the steady-state distribution from the Fokker-Planck equation. This distribution gives us information about how the properties of the stimulus and spiking dynamics determine how the system samples its subthreshold nonlinearities. We will then use the steady-state distribution to map the nonlinear models onto linear models and thus compute an approximation to the STA. Since the linear model follows from the steady-state distribution, we can think of the linear model as describing the "time-averaged dynamics" of the nonlinear models. 3.1 The steady-state distribution A key ingredient for understanding the behavior of the spiking models and for determining an analytically tractable mapping of a nonlinear model to a linear model is the steady-state probability distribution for the voltage 6 in response to an input with given statistical characteristics. This proba- bility distribution, pN (v), can be computed from the Fokker-Planck equa- tion (Paninski et al., 2003; Brunel and Latham, 2003; Lindner et al., 2003; Fourcaud-Trocme et al., 2003), which for models of the form given in equa- tion 1 is: ∂pN (v, t) ∂t = ∂ ∂v (cid:20)(cid:18) v − f (v) τm (cid:19) pN (v, t)(cid:21) + σ2 2τm ∂2pN (v, t) ∂v2 +R(t) [δ(v − Vr) − δ(v − Vs)] , (6) where R(t) is the time-dependent mean firing rate that needs to be deter- mined self-consistently in solving the equation. This is a continuity equation for pN (v, t) which expresses that the evolution of the distribution is driven by the deterministic nonlinear driving force, diffusion, and spiking. We are interested in the steady-state distribution, for which ∂pN ∂t = 0 and R(t) goes to the mean rate R. Using standard methods (Risken, 1996), one can show that the steady state distribution is: pN (v) = 2Rτm σ2 e −1 σ2 (v2 −2F (v))Z Vs max(v,Vr) dv′ e 1 σ2 (v′2 −2F (v′)), (7) stant. where F (v) = R f (v)dv, and the mean firing rate is the normalization con- This distribution is the product of a Boltzmann factor, controlled entirely by the nonlinear dynamics, F (v), preceding a spike, and a spiking flux term which carries the dependence on the spike parameters Vr and Vs. Since we are mainly interested in behavior below the unstable fixed point, or dynamical threshold, and the models considered here have reset voltages, Vr, near the resting potential, the contribution of the spiking flux term does not depend strongly on the form of F (v), but does depend strongly on the location of Vr. 3.2 Stochastic linearization We turn to a set of techniques known as stochastic linearization (SL) (see (Socha, 2005) for an extensive review) to model and understand the behavior of the STA of the nonlinear models. In the SL approach, one seeks the parameters of a linear model that optimally capture the properties of the nonlinear model in a regime of interest. In our case, we are interested in the linear model that best captures the approach of a nonlinear model to threshold for a given input standard deviation, but we are unconcerned with 7 the dynamics of the spike itself. Thus, we search for an equivalent linear model of the form: τm v = −kσv + cσ + s(t), (8) where we make no attempt to model the spike or the reset (Gerstner and Kistler, 2002). This linear model is simply an Ornstein-Uhlenbeck process (Risken, 1996), and it has the associated steady-state probability distribution: pL(v) =r kσ πσ2 exp"− kσ σ2 (cid:18)v − cσ kσ(cid:19)2# . (9) To determine the parameters of the optimal linear model, we must se- lect an optimization function that maps the nonlinear model onto the linear model. There are no unique methods for choosing optimization functions that will yield good results (Socha, 2005), and different functions will gen- erally yield different results. We focus on two optimization functions which give weight to different properties of the nonlinear model. 3.3 Minimizing the Kullback-Leibler divergence The Kullback-Leibler divergence (DKL) measures the similarity of two prob- ability distributions (Cover and Thomas, 2006). To map the nonlinear mod- els to sets of linear models, we can use the DKL to minimize the difference between the subthreshold part of the nonlinear steady-state distribution and the matched linear model's steady-state distribution. The DKL for this problem is: DKL (pN pL) = Z Vth where ZN = Z Vth −∞ −∞ dv pN (v) ZN ln pN (v) ZN pL(v) , (10) pN (v). To find the optimal linear model with this criterion, we minimize the DKL with respect to kσ and cσ. Doing so yields kσ = cσ = kσE [v] , , (11) where E [. . .] = Z −1 pN (v) [. . .] . σ2 2(cid:16)E [v2] − E [v]2(cid:17) N Z Vth −∞ 8 In the σ → 0 limit, k0 = 1 and c0 = 0, corresponding to the classical linearization around the fixed point of a nonlinear model. These expressions show that minimizing the DKL amounts to simply estimating the mean and variance below threshold. This criterion is only sensitive to the probability distribution itself and has no knowledge of the underlying dynamics. 3.4 Minimizing the energy below threshold An alternative optimization criterion is to optimize the mean square error in the energy, or first integral of the nonlinear models, below threshold. The energy of the nonlinear model is E = v2 2 − F (v), and so the optimization criterion for kσ and cσ is I = E"(cid:18) v2 2 (1 − kσ) + cv − F (v)(cid:19)2# , (12) where F (v) and E [. . .] are defined as before. This criterion amounts to trying to match the Boltzmann part of the distributions, taking spiking into account only through the bias it provides to the expectation value. This piece is primarily sensitive to the specifics of the dynamics below threshold and is less sensitive to the overall shape of the distribution than the DKL is. Minimizing I yields: kσ = 1 − 2 E(cid:2)v2F (v)(cid:3) E(cid:2)v2(cid:3) − E [vF (v)] E(cid:2)v3(cid:3) E [v4] E [v2] − E [v3]2 cσ = E [vF (v)] E(cid:2)v4(cid:3) − E(cid:2)v2F (v)(cid:3) E(cid:2)v3(cid:3) E [v4] E [v2] − E [v3]2 (13) Again, in the σ → 0 limit, k0 = 1 and c0 = 0. In this case, we see that the optimal parameters are directly sensitive to the form of the nonlinearity below threshold and that the shape of the probability distribution only enters through the expectation values. 3.5 The meanings of the optimization criteria The two optimization criteria give different weights to different roles of the nonlinearity. The DKL criterion is sensitive to the net statistical distribu- tion below threshold, regardless of whether it comes about due to the spike 9 or the subthreshold nonlinearity, whereas the energy criterion is primarily sensitive to the form of the subthreshold nonlinearity. Accordingly, we can expect that the linear model, for a given nonlinear model and input stan- dard deviation, found by the different criteria will be different. Specifically, for nonlinear models whose optimal linear equivalents are best described by the energy criterion, the parameters kσ and cσ will be closely related to the form of the subthreshold nonlinearity but may not be very sensitive to the overall details of the voltage distribution below threshold. In contrast, for models with linear equivalents that are best described by the DKL criterion, the parameters may have essentially no relation with the subthreshold non- linearity, but rather describe global statistical properties set by the mean and variance of the voltage distribution. 3.6 STA of the linear model To find the STA for the linear model and compare it to numerical simula- tions, we move to discrete time by defining t = nh, where n is an integer and h is the time step. For clarity of notation, we identify v(t) = v(nh) ≡ vn. The linear model in equation 8 is equivalently described by the forward transition probability distribution: p (vn+1vn) =r τm 2πσ2h exp"− τm 2σ2h(cid:18)vn+1 −(cid:18)1 − hkσ τm (cid:19) vn − hcσ τm (cid:19)2# . (14) Also of use are the steady-state probability distribution, pL(v), given in equation 9, and the backward transition probability distribution, p(vnvn+1), which can be derived with Bayes' rule: p (vnvn+1) = p(vn+1vn)pL(vn) pL(vn+1) , =r τm 2πσ2h exp"− τm 2σ2h(cid:18)vn −(cid:18)1 − hkσ τm (cid:19) vn+1 − hcσ τm (cid:19)2# , (15) for small h . Notice that the linear model is statistically reversible (Weiss, τm 1975): the backward transition distribution is the time reversal of the for- ward, vn ⇄ vn+1. Since the model is linear, the STA follows from the spike- triggered voltage, ¯v, via: ¯s(t) = τm ¯v(t) + kσ ¯v(t) − cσ, ¯sn = (¯vn − ¯vn−1) 10 τm h + kσ ¯vn−1 − cσ. (16) (17) The spike-triggered voltages of the linear model can be found exactly with the following recipe. We start at the spike time, t = 0 (n = 0) -- the first time for which v > Vth and v > 0. The mean voltage at the spike time, ¯v0, is given by: ¯v0 =Z ∞ −∞ dv0 v0 p(v0spike) (18) The spike-triggered voltage distribution, p(v0spike), follows from the threshold- crossing condition. The probability of finding a voltage v0 at the spike time is given by the probability that v0 is above Vth, multiplied by the probability that v0 was arrived at from voltages v−1 that were below threshold, summed over all possible subthreshold values of v−1: p(v0spike) = Z −1 0 H(v0 − Vth)Z Vth −∞ dv−1p(v0v−1)p(v−1), (19) where p(v−1) is the unconditioned distribution of voltages prior to the spike and is given by the steady state distribution in equation 9, p(v0v−1) is the forward transition distribution, H(v0 − Vth) is the Heaviside function representing the probability for v0 to be above threshold, and Z0 is the normalization constant. The mean voltage at the time immediately preceding the spike, ¯v−1, is determined by averaging over all voltages below threshold that can transition to voltages above threshold in the next time step, and can be found from ¯v−1 = Z ∞ −∞ dv−1 v−1 p(v−1spike), (20) p(v−1spike) = Z −1H(Vth − v−1)Z ∞ Vth dv0 p(v−1v0)p(v0spike), (21) where Z is the normalization constant for this distribution. Similarly, the remaining ¯vn for n ≤ −2 are given by: ¯vn =Z ∞ −∞ dvn . . . dv−1 vn p(vnvn+1) . . . p(v−1spike). (22) Equation 22 is exact for a linear model but is impractical to use. With- out noticeable loss of accuracy for the simulations considered in this paper, numerous simplifications can be made. For a discussion of approximations to ¯v0 and ¯v−1, see the appendix. 11 Via the central limit theorem, the ¯vn for n ≤ −2 can be simplified as: ¯vn ≈Z ∞ −∞ dvn vn p(vn¯vn+1). (23) Since these are all gaussian integrals over an infinite domain, the mean value is the most likely value and so the remaining averages are arrived at recursively to give: ¯vn =(cid:18)1 − hkσ τ (cid:19) ¯vn+1 + hcσ τ for n ≤ −2. (24) Using the fact that h of the exponential function and gives: τm ≪ 1, this recursion relation can be solved in terms ¯vn = cσ kσ +(cid:18)¯v−1 − cσ kσ(cid:19) exp(cid:20) kσ(n + 1)h τm for n ≤ −1. (25) (cid:21) The STA immediately follows from equation 16, and is: ¯sn =(2kσ(cid:16)¯v−1 − cσ kσ(cid:17) exph kσ(n+1)h kσ ¯v0 + (¯v0 − ¯v−1) τm h τm i n ≤ −1, n = 0. (26) For the linear model, the STA is simply given by the exponential filter up to a singular piece at the spike time that arises from requiring that threshold be crossed from below. 3.7 Comparison to numerics For the QIF, the energy criterion qualitatively captures the time constant at all σ and correctly matches the amplitude of the STA at high σ (see figures 2A and 3C). Conversely, the DKL criterion is much less accurate. While it too predicts qualitatively correct time constants, the amplitude of the predicted STAs is much too large (not shown). This is because the DKL criterion strongly weights the effects of the after-spike reset and total probability mass, and thus biases the resting potential of the linear model too far below threshold. Thus, the STA for the QIF is primarily determined by the form of the subthreshold nonlinearity and is less sensitive to the escape from the subthreshold domain due to spiking. Figure 3A shows how the optimal linear model relates to the subthreshold nonlinearity in this case. For the EIF, however, the energy criterion gives kσ ≈ 1 for all σ. While this is not surprising given the effectively linear subthreshold dynamics, it 12 does not agree with the numerics. The DKL criterion, on the other hand, applied to the EIF leads to qualitative agreement between the linear mod- els and numerics (see figures 2B and 3D). This confirms the idea that the changes in the STA in the EIF can only be due to the reduction in the time spent below threshold because of spiking, and that the small amount of non- linearity below threshold for the parameters used is not relevant except at small σ (see figure 3B for further discussion). These changes in the STA are analogous to those studied previously by Paninski (Paninski et al., 2005). The singular upward step at the spike time arises from the condition that threshold must be crossed from below -- that v must be positive -- to elicit a spike (Aguera y Arcas et al., 2003; Hong et al., 2007). This "delta-function" component, shown in figures 1C and 1D, appears here so prominently be- cause we have chosen the spike-defining threshold to be at a voltage for which the stimulus is still relevant to spiking. This step does not vanish in the continuous-time limit. The value of the step can be calculated ap- proximately (see appendix) with good agreement with simulation data (see figure 4). This mode occasionally appears in experimental STAs when the spike waveform is slow (R. Mease, personal communication). It is usually not seen because the threshold is generally drawn well into the intrinsically excitable domain of the voltage and so a condition on v does not significantly constrain the stimulus in that situation. 4 Discussion Due to nonlinearity, LN characterizations of neural systems show depen- dence on stimulus statistics, even without changes in the underlying dy- namical parameters (Theunissen et al., 2000; Yu and Lee, 2003; Borst et al., 2005; Gaudry and Reinagel, 2007; Gill et al., 2008; Westwick and Kearney, 2003). In particular, by changing only the input standard deviation, the effective computation changes its functional form and timescale. We have explored the consequences of this for two reduced naturally-spiking neu- ron models, the quadratic and exponential integrate-and-fire models. In determining the linear filter or filters characterizing the model, our work differs significantly from previous approaches (Gerstner and Kistler, 2002; Hong et al., 2007; Wilson and Steyn-Ross, 2008; Badel et al., 2008a) in that the point of linearization is not taken, as is classically done, to be the equilibrium point; rather, we allow the subthreshold voltage distribution to determine the optimal point of linearization. This distribution carries information about the form of the nonlinearities, the mean firing rate, and 13 the stimulus itself. These properties account for changes in the effective linear model with stimulus variance. We find that despite these models' su- perficial similarities, different mechanisms are primarily responsible for this form of adaptation. The key difference between the models is that the QIF is nonlinear below threshold, qualitatively corresponding to a neuron with hyperpolarizing currents that are activated below threshold, while the EIF is mostly linear below threshold. Both models have been successfully fit to neuronal data from a variety of neuron types (Izhikevich, 2004; Rauch et al., 2003; Badel et al., 2008b). Thus, both the neuron's intrinsic properties and the statistics of the background or of the driving stimulus ensemble determine the effective fil- tering properties of the system. This shows one means by which modulating the statistics of the input can effectively gate the transmission of different types of input or stimulus features through the system (Hasenstaub et al., 2007; Destexhe and Pare, 1999; Fellous et al., 2003). While this analysis fo- cused on very simple model neurons, the methods we describe generalize to more complex, higher dimensional neuronal models, although analytical so- lutions are unlikely. These simple examples give a clear insight into intrinsic modulation of feature selectivity. Our previous treatments of this problem (Aguera y Arcas et al., 2003; Aguera y Arcas and Fairhall, 2003; Hong et al., 2007) concentrated on the case where spikes are well-separated, so that the effects of spike history are explicitly separated from the role of the stimulus in determining the proba- bility of generating a spike. Another approach to this problem is to include an explicit spike-history term in the generative model (Gerstner and Kistler, 2002; Paninski et al., 2004; Powers et al., 2005). Here, the spike history is incorporated into the computation of features due to the effects of the mean firing rate on the steady-state distribution of threshold escape and reset. These results underscore the difficulty in inferring information about underlying biophysical parameters from the output of reverse correlation, independent of a consideration of the stimulus properties. 5 Acknowledgments We thank Brian Lundstrom, Sungho Hong, Liam Paninski, and our review- ers for helpful comments and discussions. This research is funded by the McKnight Endowment Fund for Neuroscience. 14 Appendix: Approximating the singular piece of the STA In numerical investigation, we find that there is a simple approximation for the average voltage at the spike time, ¯v0, for the range of σ considered in this paper and the use of the dynamical threshold for spike triggering. In our hands, this relationship does not seem to follow from an obvious perturbative calculation. For the QIF and EIF, we find to first order in σq h τm : (27) ¯v0 ≈ Vth − cσ kσ + f σr h τm where f = 0.85 is the result of a fit to the exact integral in equation 18 for different σ, h, and τm. The exact integral for the average voltage at the time immediately pre- ceding the spike, ¯v−1, can also be approximated with an error of a few parts in a thousand. The distribution, p(v−1spike), can be approximated as: p(v−1spike) ≈ H(Vth − v−1)p(v−1¯v0) Z Vth −∞ dv−1p(v−1¯v0) (28) To first order in σq h show that: τm , where ¯v0 is given by equation 27, it is possible to ¯v−1 ≈ Vth − cσ kσ − σr h τm r 2 π + f(cid:18) 2 π − 1(cid:19)! (29) Taken together, this shows that the singularity in the STA at the spike h . See figure 4 for numerical results. time, which is given by v, goes as σp τm References Abeles, M. (1982). Role of the cortical neuron: integrator or coincidence detector? Israel J. Med. Sci., 18:83 -- 92. Aguera y Arcas, B. and Fairhall, A. L. (2003). What causes a neuron to spike? Neural Computation, 15:1789 -- 1807. Aguera y Arcas, B., Fairhall, A. L., and Bialek, W. (2003). Computation in a single neuron: Hodgkin and Huxley revisited. Neural Computation, 15:1715 -- 1749. 15 Badel, L., Gerstner, W., and Richardson, M. J. (2008a). Spike-triggered averages for passive and resonant neurons receiving filtered excitatory and inhibitory synaptic drive. Phys. Rev. E, 78(011914). Badel, L., Lefort, S., Berger, T., Petersen, C., Gerstner, W., and Richard- son, M. (2008b). Extracting non-linear integrate-and-fire models from experimental data using dynamic i-v curves. Biological Cybernetics, 99(4 -- 5):361 -- 370. Borst, A., Flannigan, V. L., and Sompolinksy, H. (2005). Adaptation with- out parameter change: Dynamic gain control in motion detection. PNAS, 102(17):6172 -- 6176. Brunel, N. and Latham, P. E. (2003). Firing rate of the noisy quadratic integrate-and-fire neuron. Neural Computation, 15:2281 -- 2306. Bryant, H. and Segundo, J. (1976). Spike initiation by transmembrane current: a white-noise analysis. J. Physiol., 260:279 -- 314. Cover, T. M. and Thomas, J. A. (2006). Elements of Information Theory. Wiley. de Boer, E. and Kuyper, P. (1968). Triggered correlation. IEEE Trans. Biomed. Engr., 15:169 -- 179. de Ruyter van Steveninck, R. and Bialek, W. (1988). Real-time performance of a movement sensitive neuron in the blowfly visual system: Coding and information transfer in short spike sequences. Proc. R. Soc. London Ser. B, 234:379 -- 414. Destexhe, A. and Pare, D. (1999). Impact of network activity on the integra- tive properties of neocortical pyramidal neurons in vivo. J. Neurophysiol., 81:1531 -- 1547. Ermentrout, G. B. and Kopell, N. (1986). Parabolic bursting in an excitable system coupled with slow oscillation. SIAM J. Appl. Math., 46:233 -- 253. Fellous, J., Rudolph, M., Destexhe, A., and Sejnowski, T. (2003). Synaptic background noise controls the input/output characteristics of single cells in an in vitro model of in vivo activity. Neuroscience, 122:811 -- 829. Fourcaud-Trocme, N., Hansel, D., van Vreeswijk, C., and Brunel, N. (2003). How spike generation mechanisms determine the neuronal response to fluctuating inputs. J. Neurosci., 23(37):11628 -- 11640. 16 Gaudry, K. S. and Reinagel, P. (2007). Contrast adaptation in a nonadapting LGN model. J. Neurophysiol., 98:1287 -- 1296. Gerstner, W. and Kistler, W. M. (2002). Spiking Neuron Models: Sin- gle Neurons, Populations, Plasticity. Cambridge University Press, Cam- bridge. Gill, P., Zhang, J., Woolley, S., Fremouw, T., and Theunissen, F. (2008). Sound representation methods for spectro-temporal receptive field esti- mation. J Comp. Neurosci., 21(1):5 -- 20. Hasenstaub, A., Sachdev, R., and McCormick, D. (2007). State changes rapidly modulate cortical neuronal responsiveness. J. Neurosci., 27:9607 -- 9622. Hong, S., Aguera y Arcas, B., and Fairhall, A. L. (2007). Single neuron computation: from dynamical system to feature detector. Neural Com- putation, 19:3133 -- 3172. Hunter, I. and Korenberg, M. (1986). The identification of nonlinear bi- ological systems: Wiener and hammerstein cascade models. Biological Cybernetics, 55:135 -- 144. Izhikevich, E. (2000). Neural excitability, spiking, and bursting. Interna- tional Journal of Bifurcation and Chaos, 10:1171 -- 1266. Izhikevich, E. (2001). Resonate-and-fire neurons. Neural Networks, 14:883 -- 894. Izhikevich, E. (2004). Which model to use for cortical spiking neurons? Neural Networks, IEEE Transactions on, 15(5):1063 -- 1070. Lindner, B., Longtin, A., and Bulsara, A. (2003). Analytic expressions for rate and CV of a type I neuron driven by gaussian white noise. Neural Computation, 15:1761 -- 1788. Meister, M. and Berry II, M. (1999). The neural code of the retina. Neuron, 22:435 -- 450. Paninski, L. (2006a). The most likely voltage path and large deviations approximations for integrate-and-fire neurons. J. Comp. Neurosci., 21:71 -- 87. Paninski, L. (2006b). The spike-triggered average of the integrate-and-fire cell driven by gaussian white noise. Neural Computation, 18:2592 -- 2616. 17 Paninski, L., Lau, B., and Reyes, A. (2003). Noise-driven adaptation: in vitro and mathematical analysis. Neurocomputing, 52-54:877 -- 883. Paninski, L., Pillow, J., and Simoncelli, E. (2004). Maximum likelihood es- timation of a stochastic integrate-and-fire neural encoding model. Neural Computation, 16:2533 -- 2561. Paninski, L., Pillow, J., and Simoncelli, E. (2005). Comparing integrate- and-fire models estimated using intracellular and extracellular data. Neu- rocomputing, 65-66:379 -- 385. Computational Neuroscience: Trends in Research 2005. Powers, R., Dai, Y., Bell, B., Percival, D., and Binder, M. (2005). Contri- butions of the input signal and prior activation history to the discharge behaviour of rat motoneurones. J. Physiol., 528:131 -- 150. Prescott, S. A., Ratte, S., De Koninck, Y., and Sejnowski, T. J. (2006). Nonlinear interaction between shunting and adaptation controls a switch between integration and coincidence detection in pyramidal neurons. J. Neurosci., 26(36):9084 -- 9097. Rauch, A., Camera, G. L., Luscher, H., Senn, W., and Fusi, S. (2003). Neocortical pyramidal cells respond as integrate-and-fire neurons to in vivo-like input currents. J. Neurophysiol., 90(3):1598 -- 1612. Rieke, F., Warland, D., de Ruyter van Steveninck, R., and Bialek, W. (1996). Spikes: exploring the neural code. The MIT Press, Cambridge, MA. Risken, H. (1996). The Fokker-Planck Equation: Methods of Solution and Applications. Springer, Berlin, second edition. Sakai, H. (1992). White-noise analysis in neurophysiology. Physiol. Rev., 72:491 -- 505. Simoncelli, E., Paninski, L., Pillow, J., and Schwartz, O. (2004). Character- ization of neural responses with stochastic stimuli., pages 327 -- 338. The MIT Press, Cambridge, MA, third edition. Slee, S., Higgs, M., Fairhall, A., and Spain, W. (2005). Two-dimensional time coding in the auditory brainstem. J. Neurosci., 26:9978 -- 9988. Socha, L. (2005). Linearization in analysis of nonlinear stochastic systems: Recent results -- part I: Theory. Applied Mechanics Reviews, 58(3):178 -- 205. 18 Svirskis, G., Dodla, R., and Rinzel, J. (2003). Subthreshold outward currents enhance temporal integration in auditory neurons. Biol. Cybern., 89:330 -- 340. Svirskis, G., Kotak, V., Sanes, D., and Rinzel, J. (2004). Sodium along with low threshold potassium currents enhance coincidence detection of subthreshold noisy signals in mso neurons. J. Neurophysiol., 91:2465 -- 2473. Theunissen, F. E., Sen, K., and Doupe, A. J. (2000). Spectral-Temporal Receptive Fields of Nonlinear Auditory Neurons Obtained Using Natural Sounds. J. Neurosci., 20(6):2315 -- 2331. Weiss, G. (1975). Time-reversibility of linear stochastic processes. J. Appl. Probab., 12(4):pp. 831 -- 836. Westwick, D. T. and Kearney, R. E. (2003). Identification of Nonlinear Phys- IEEE Press Series in Biomedical Engineering. IEEE iological Systems. Press. Wilson, M. T. and Steyn-Ross, D. A. (2008). Subthreshold dynamics of a single neuron from a Hamilonian perspective. Phys. Rev. E, 78(061908). Yu, Y. and Lee, T. S. (2003). Dynamical mechanisms underlying contrast gain control in single neurons. Phys. Rev. E, 68(1):011901. 19 Figure 1: The STAs for (A) the QIF and (B) the EIF (using the L2-norm for easy visual comparison), triggered on Vth = α−1, for various σ with the upward step at t = 0 removed. Note that at small σ, the STA is non- monotonic while, at large σ, it is approximately a decaying exponential. As representative examples, the STA in real units is shown (C) for the QIF for σ = 0.3 and (D) for the EIF for σ = 3 with the last time-step included. 20 Figure 2: Comparison between the numerical STA (solid) and the STA predicted by equation 26 (dashed). (A) For the QIF, numerical results (solid) are compared to the prediction from stochastic linearization via the Energy criterion (dashed). In this case, the DKL criterion predicts STA amplitudes that are too large (not shown) (B) For the EIF, numerical results (solid) are compared to the prediction from SL via the DKL criterion. In this case, the Energy criterion predicts kσ ≈ 1 for all σ (not shown). 21 Figure 3: The left column refers to the QIF and its optimization via the energy criterion, and the right, the EIF via the DKL. (A) This figure shows how the linear model corresponding to the energy criterion for σ = 2 for the QIF corresponds to the full nonlinear model. In gray, we see the steady- state voltage distribution. The quadratic nonlinearity is shown dashed and the optimal linear model as determined by the energy criterion is the solid line. We see that the linear model in this case is closely related to the average slope of the quadratic nonlinearity below threshold. (B) In contrast, for the EIF, using the DKL criterion, the optimal linear model does not correspond closely with the exponential nonlinearity. This is indicative of the fact that the adaptation of the STA in the EIF is due primarily to the spiking reset, as evinced by the STA results (see figure 2). (C,D) The steady state distributions of the (QIF,EIF) models (solid lines) are compared to their linear model approximations (dashed). 22 Figure 4: The value of the STA in the singular component at the time of the spike, ¯s(0), for all cases studied in this paper. As explained in the appendix, the value is approximately linear in σ and model-independent. 23
1804.01958
1
1804
2018-04-05T17:12:17
The multiphysics of prion-like diseases: progression and atrophy
[ "q-bio.NC", "q-bio.TO" ]
Many neurodegenerative diseases are related to the propagation and accumulation of toxic proteins throughout the brain. The lesions created by aggregates of these toxic proteins further lead to cell death and accelerated tissue atrophy. A striking feature of some of these diseases is their characteristic pattern and evolution, leading to well-codified disease stages visible to neuropathology and associated with various cognitive deficits and pathologies. Here, we simulate the anisotropic propagation and accumulation of toxic proteins in full brain geometry. We show that the same model with different initial seeding zones reproduces the characteristic evolution of different prion-like diseases. We also recover the expected evolution of the total toxic protein load. Finally, we couple our transport model to a mechanical atrophy model to obtain the typical degeneration patterns found in neurodegenerative diseases.
q-bio.NC
q-bio
The multiphysics of prion-like diseases: progression and atrophy †Stevens Institute of technology, New Jersey, USA Johannes Weickenmeier†, Ellen Kuhl‡, and Alain Goriely∗ ‡Living Matter Laboratory, Stanford University, Stanford USA, ∗Mathematical Institute, University of Oxford, Oxford UK . Many neurodegenerative diseases are related to the propagation and accumulation of toxic proteins throughout the brain. The lesions created by aggregates of these toxic proteins further lead to cell death and accelerated tissue atrophy. A striking feature of some of these diseases is their characteristic pattern and evolution, leading to well-codified disease stages visible to neuropathology and associated with various cognitive deficits and pathologies. Here, we simulate the anisotropic propagation and accumulation of toxic proteins in full brain geometry. We show that the same model with different initial seeding zones reproduces the characteristic evolution of different prion- like diseases. We also recover the expected evolution of the total toxic protein load. Finally, we couple our transport model to a mechanical atrophy model to obtain the typical degeneration patterns found in neurodegenerative diseases. PACS numbers: Introduction: Age-related neurodegenerative disorders are extremely complex and multifaceted pathologies. Yet, their evolution is known to be closely associated with the progression of particular protein aggregates. When these proteins misfold and/or aggregate they can be transported into the brain tissue and become the seed for further misfolding and aggregation [1]. Unless these toxic proteins are removed or stabilized, this chain- reaction proceeds. As toxic proteins invade the brain and form larger aggregates, they prevent the proper function of other proteins and create tissue lesions. over time, it leads to a disruption of the proper function of the nervous system, a loss of tissue structure, necrosis, brain atrophy, and ultimately death [2]. In prion diseases, such as Creutzfeldt-Jakob disease, the infectious agents are typically small soluble misfolded isoforms of proteins that can aggregate and trigger the misfolding of the same protein found in its benign form. When these aggregates are large enough, they form char- acteristic tissue lesions. A similar mechanism of pro- tein corruption and propagation is found in both sec- ondary injury following traumatic brain injury [3–5] and common age-related neurodegenerative diseases such as Alzheimer's disease, the most common form of demen- tia [6]. For instance, it is known from in vitro exper- iments that the amyloid-β protein can form aggregates that propagate through the tissues. These aggregates have a distribution of sizes from small soluble groups to large insoluble fibrils. Their accumulation as plaques is believed to play a key role in Alzheimer's disease. Tau proteins also play an important role in these diseases [7]. These small proteins are known to stabilize mi- crotubules within the axon. Hyperphosphorylated tau proteins can act as seeds for further misfolding and ag- gregation. These aggregates are found within the axon where they are rapidly transported through the usual interneuronal transport mechanisms [17]. They are also transported from the cell to the extracellular space where they diffuse through secretion and damage of the host cell, and from the extracellular space into cells by endo- cytosis [8, 9]. These aggregates are subject to biological clearance slowing down or arresting the overall spreading process. The particular role of neuronal pathways in this transport mechanism may provide an explanation for the observation that the spreading of neurofibrillary tangles follows a characteristic topographic pattern. FIG. 1: Typical spatial progression of protein aggregates in various neurodegenerative diseases. From Jucker [1]. Independently of their molecular origin or specific ac- tions, these pathologies, known as prion-like diseases, share the same macroscopic features when viewed as a a. amyloid-β in Alzheimer's diseaseb. tau inclusions in Alzheimer's diseasec. α−synuclein inclusions in Parkinson's diseased. TPD-43 inclusion in amyotrophic lateral sclerosis spatio-temporal evolution process [10]. The progression is believed to include the following steps: (i) seeding of misfolded proteins, (ii) templated misfolding and aggre- gation of native (homologous) proteins, (iii) growth of aggregates, (iv) fragmentation and spatial propagation of aggregates of different sizes, (v) assembly of misfolded proteins into secondary structures (e.g. protofibrils and fibrils), and (vi) formation of larger tertiary structures (lesions) at the tissue level. A striking feature of each par- ticular pathology is its systematic evolution pattern. We posit that the main reason for such reproducibility is that the overall progression of the disease is governed by basic transport processes and that the key difference between individual diseases, aside from the particular set of pro- teins involved, is the original location of the seeds, lead- ing to a characteristic spatio-temporal evolution (Fig. 1) with respective pathological symptoms. An essential feature of the dynamics is that small proteins are preferentially transported transynaptically but can also diffuse from the extracellular space to the cells through endocytosis [11]. These effects lead to fast transport along neuronal pathways and a slower diffusion in directions perpendicular to the axon bundles. Here, we couple an anisotropic transport process with a mechanical model to take into account the relative effect of seed location, anisotropy, and convoluted brain geometry on the damage patterns and the resulting atrophy of brain tissue [12, 13]. We show that a minimal model for transport and atrophy is sufficient to recover the main qualitative spatial features of different neurodegenerative diseases as well as the overall increase of key biomarkers. The model: We couple a transport model to a mechani- cal model of the brain. The geometry of the brain is taken to be a connected domain Ω in either R2 (slice simulation) or R3 (full brain simulation) composed of gray (Ωgray) and white (Ωwhite) matters with different material prop- erties and total volume (or area) V . We first consider the transport process. Different models for the aggre- gation/propagation process have been proposed ranging from graph Laplacian diffusion on simplified structural networks [14] to Smoluchowski-type equations [15, 16] tracking the aggregate-size distribution. Here, we use a minimal model for the propagation of a toxic protein that captures the important physical characteristics of the problem: transformation of existing native proteins and anisotropic diffusion along axonal pathways. We as- sume that the concentration c(x, t) ∈ Ωwhite follows the Fisher-KPP equation ∂c ∂t = ∇ · (D∇c) + αc(1 − c), (1) where D = d⊥1 + (d(cid:107) − d⊥)γ ⊗ γ is a transversely anisotropic diffusion tensor chosen to have a preferential direction along the axon bundle characterized by the unit 2 vector γ = γ(x, t). Here, α > 0, d⊥ is the regular tissue diffusion and d(cid:107) (cid:29) d⊥ is the diffusion along the axons. This model describes the overall increase of toxic protein concentration, assuming that the pool of protein in the native form is sufficiently large as not to be affected by the increase in toxic protein. The nonlinearity provides a saturation term expressing the maximal concentration of toxic proteins achievable (taken to be 1 without loss of generality). In one dimension (along the axonal bun- dles), the system supports traveling fronts with a velocity v = 2(cid:112)αd(cid:107) leading to a complete invasion in a charac- teristic time T = L/v, where L is the typical length scale for the entire brain, which is on the order of 16cm for hu- man brains. In the gray matter, for x ∈gray, we assume simple isotropic diffusion from cells to cells different from the white matter, (Eq. (1) with d(cid:107) = d⊥ and α = 0). We are interested in tracking two quantities related to the concentration field c(x, t). First, the time τ (x, Ccrit) at which the concentration first reaches a critical level close to the saturation level Ccrit = (1 − ε) defined im- plicitly by c(x, τ ) = Ccrit. Second, we investigate the temporal evolution of the average amount of toxic pro- tein as a possible biomarker, computed as (cid:90) Ω C(t) = 1 V c(x, t) dx. (2) We couple the transport process to the mechanics of atrophy by assuming that gray and white matter tissues are morphoelastic materials [18, 19] and assigning an isotropic shrinking factor 0 < ϑ < 1 at each point depending on the concentration of toxic proteins. Given an initial reference configuration, atrophy is character- ized by a deformation ϕ : B0 → Bt, x(X, t) = ϕ(X, t) from the reference brain B0 to the aged brain after shrinking Bt. Here, x(X, t) is the position at time t of the material point originally located at X at time t = 0. This deformation is obtained by first defining the deformation gradient F = ∇Xϕ and assuming that it can be decomposed as F = ϑ1/3A, where ϑ denotes the volumetric change (ϑ < 1 represents shrinking and ϑ > 1 would correspond to growth). The elastic defor- mation tensor A, is computed by solving the Cauchy equation div(T) = 0 for the Cauchy stress tensor T = J−1A · ∂AW associated with the strain-energy density W = W (A, x) and J = det(A). The amount of shrinking is proportional to the total exposure of the tissue to the toxic proteins, so that ∂tϑ = δc, where δ > 0 is an overall parameter that describes the rate of removal of material. Computational investigation: We create three finite element models from T2-weighted magnetic resonance images of a 32-year old male using Simpleware: a two- dimensional sagittal model, a two-dimensional coronal model, and a fully three-dimensional, anatomically ac- curate whole brain model. The sagittal model consists of 13442 linear triangular elements and 7216 nodes; the coronal model consists of 38223 linear triangular elements and 19813 nodes; and the three-dimensional whole brain model consists of 401940 linear tetrahedral elements and 80233 nodes. To model anisotropic toxic protein propa- gation, we register the axonal fiber orientation from the diffusion tensor magnetic resonance image of the subject and create an element based fiber-orientation map. For the initial seeding, we assume a concentration c(x) = 1 for all x ∈ Ωseed ⊂ Ω and c(x) = 0 other- wise. Then, using Eq. (1) we propagate the toxic protein in space and time across the brain with a diffusivity con- trast of d(cid:107) : d⊥ = 100 to ensure a faster propagation along the axonal fiber direction. In the white matter, we use d(cid:107) = 100 mm2/year and α = 0.5/year. In the gray matter, we use d(cid:107) = d⊥ = 10 mm2/year. We post-process the concentration field c to extract the activation time τ (x, Ccrit). From those activation times, we create activation maps of all four cases of prion-like disease similar to Fig. 1. Using Eq. (2), from the spatio- temporal concentration field c(x, t), we extract the inte- gration of the toxic protein concentration in time C(t) to create a temporal map of the toxic protein concentration averaged over the entire brain [20]. FIG. 2: Activation maps of tau inclusions in brains of patients with Alzheimer's disease in a 2D sagittal slice. Finally, using the tissue atrophy model with different atrophy rates δ in the gray and white matter tissue, we utilize our two-dimensional coronal model to create atrophy maps. It is important to note that atrophy takes place primarily in locations with elevated con- centrations of toxic proteins. Therefore, the change of volume does not significantly affect the time evolution of the concentration since it is close to maximal at the locations where atrophy takes place. We take advantage of this property to compute tissue shrinking 3 as a post-processing step based on the values of the concentration at different time points. Since we assume that after each step of shrinking, there is total relaxation of the stresses, the mechanical stresses generated are small during shrinking and we can use a standard compressible neo-Hookean model valid in small defor- mations with parameters fitted to experimental data: W = c1(tr(AT · A)− 3) + c2(J − 1)2 with c1 = 666.56 Pa and c2 = 1777.5 Pa [21]. For the relative shrinking rate, it is known that the gray matter tissue shrinks looses more volume than the white matter tissues [22]. For our computation, we use a ratio ϑgray : ϑwhite = 4 : 1. 2D simulations: We start with a two-dimensional simulation on a sagittal slice (Fig. 2). The region in- cludes the ventricle and cerebellum but, through no-flux boundary conditions, it is assumed that no progression takes place into these regions. Based on the description of the evolution of tau inclusions in Alzheimer's disease, we simulate the propagation of toxic tau proteins by considering an initial seeding close to the entorhinal cor- tex (Fig. 2a). We observe the quick progression into the brain stem and the hippocampus (Fig. 2a), to paralimbic and adjacent medial-basal temporal cortex (Fig. 2b), to cortical association areas (Fig. 2c), and eventually reaching primary sensory-motor and visual areas. The relatively rapid progression along the hippocampus is associated with a strong anisotropy along the axonal pathways tangent to the ventricles. This progression is consistent with the well-established Braak stages of Alzheimer's disease [23]: Stage I-II transentorhinal (Fig. 2a); Stage III-IV limbic (Fig. 2b); Stage V-VI isocortical (Fig. 2c-d). 3D simulations: The previous analysis uses a high- resolution mesh but lacks the three-dimensional geom- etry of the brain. Next, we use the accepted knowledge about different neurodegenerative diseases to compare the progression of various toxic proteins. To test the specific role of geometry in the progression, we assign the same parameter values for all cases and only modify the initial seeding region of the four cases discussed in [1] and depicted in Fig. 1. For every individual seeding region we show, in Fig. 3, the time τ to reach a critical level of damage. We observe generic progression trends in all cases: (i) as soon as a sufficient level of toxic proteins reaches the area close to the ventricles, there is a fast progression in the limbic system (around the ventricles) leading to a rapid invasion of the temporal and occipital lobes; (ii) once a sufficient level of toxic protein is reached within the cortex, further invasion through the cortex takes place; (iii) if the parietal lobe is not directly involved initially, it only becomes invaded in the last stages of the disease. Tau inclusions in brains of patients with Alzheimer's diseasettinitial seedingcerebelluma.b.c.d.ventricle 4 [20]. The variation of the typical parameters entering the sigmoids are used to represent physiological quan- tities such as time of outbreak, and rate of spreading. These curves serve as a general guideline to understand the variations between individuals, pathologies, and the effect of various stimuli or therapeutics [25]. We can extract these curves from the simulation using (2) as shown in Fig. 4. Our simulations recover the predicted behavior for this type of disease integrated across the entire brain. This profile is not unexpected. Indeed, we known that Equation (1) supports fronts in one-dimension that have profile and similar evolution equations for Alzheimer's diseases. A simple estimate using the typical wave speed of front propagation in one-dimension leads to a lower and upper bound for the total invasion of 11.3 and 35.8 years, respectively, obtained by using the front velocity with either the fast axonal diffusion constant d(cid:107) ≈ 100 mm2/year or the slow gray matter diffusion constant d ≈ 10 mm2/year. These estimates are consistent with the characteristic time scales observed in Fig. 4. Similar curves have also been obrained on 2D slices in different geometries [16]. Here, we observe that the fully 3D evolution on an actual brain model also follows the same general trend of a progressive invasion. Shrinking: We now turn to the effect of elevated toxic proteins on the tissue. It is known that the formation of large aggregates or high concentration of toxic proteins prevents the proper function of neuronal cells, leading to ischemia, and eventually tissue removal. For instance, in the top row of Fig. 5, there is a marked and rapid atro- phy associated with the disease. We use our propagation model on a similar coronal slice to compute the activa- tion time at three different time points (middle row). For each of these time points we compute the shrinking of the slice and show the change in geometry from the previous one (bottom row). The degeneration-induced atrophy patterns (bottom row) agree well with the at- rophy pattern observed in Alzheimer's disease (top row). Conclusions: Neurodegenerative diseases are known for their extreme complexity and medical science has wres- tled with this major challenge by amassing a consider- able amount of information despite few successes. One of the remarkable features that appear in these diseases is their remarkable reproducible topographic propaga- tion pattern. Each disease has a well-defined spatio- temporal evolution that has been documented over the years and catalogued into stages, each associated with typical symptoms. This evolution appears much more controlled than other diseases such as cancer and has led to three main hypotheses in recent years. First, it has been proposed that diseases share the same charac- teristics as prion diseases and are based on the seeding, propagation, and accumulation of toxic proteins involved FIG. 3: Simulated activation maps for the spatial progres- sion of toxic protein in full brain geometry for various initial seeding regions matching Fig. 1 [1]. FIG. 4: Evolution of the total concentration for the simulation of Fig 2 (2D sagittal slice) and of Fig 3b (3D tau concentration in Alzheimer's disease). An indirect way to follow the evolution of neurodegen- erative diseases is to look at the averaged concentration of toxic proteins or associated biomarkers. The paradigm for this evolution is based on the so-called Jack's curves. These curves are the expected, mostly hypothetical, evolution of a typical biomarker concentration as a func- tion of time [24]. They take the shape of sigmoid-like functions: a slow evolution at first during incubation, followed by a sharp increase during the outbreak and leading to saturation when the disease is fully established a. amyloid-β in Alzheimer's diseaseb. tau inclusions in Alzheimer's diseasec. α−synuclein inclusions in Parkinson's diseasettttd. TPD-43 inclusion in amyotrophic lateral sclerosis0102030400.00.20.40.60.81.0onset2Dslice3DbrainC(t)C(t)t (years) 5 [3] A. C. McKee, T. D. Stein, C. J. Nowinski, R. A. Stern, D. H. Daneshvar, V. E. Alvarez, H.-S. Lee, G. Hall, S. M. Wojtowicz, C. M. Baugh, et al., Brain 136, 43 (2013). [4] D. H. Daneshvar, L. E. Goldstein, P. T. Kiernan, T. D. Stein, and A. C. McKee, Molecular and Cellular Neuro- science 66, 81 (2015). [5] M. Cruz-Haces, J. Tang, G. Acosta, J. Fernandez, and R. Shi, Translational neurodegeneration 6, 20 (2017). [6] J. C. Watts, C. Condello, J. Stohr, A. Oehler, J. Lee, S. J. DeArmond, L. Lannfelt, M. Ingelsson, K. Giles, and S. B. Prusiner, Proceedings of the National Academy of Sciences 111, 10323 (2014). [7] M. Goedert, Science 349, 1255555 (2015). [8] C. Soto, Neuron 73, 621 (2012). [9] J. W. Wu, M. Herman, L. Liu, S. Simoes, C. M. Acker, H. Figueroa, J. I. Steinberg, M. Margittai, R. Kayed, C. Zurzolo, et al., Journal of Biological Chemistry 288, 1856 (2013). [10] L. C. Walker and M. Jucker, Annual review of neuro- science 38, 87 (2015). [11] B. Frost, R. L. Jacks, and M. I. Diamond, Journal of Biological Chemistry 284, 12845 (2009). [12] A. Goriely, M. G. D. Geers, G. A. Holzapfel, J. Jayamo- han, A. J´erusalem, S. Sivaloganathan, W. Squier, J. A. W. van Dommelen, S. L. Waters, and E. Kuhl, Biomech. Model. Mechanobiol. 14, 931 (2015). [13] A. Goriely, S. Budday, and E. Kuhl, Advances in Applied Mechanics 48, 79 (2015). [14] F. Abdelnour, H. U. Voss, and A. Raj, Neuroimage 90, 335 (2014). [15] P. C. Bressloff, Lecture Notes on Mathematical Mod- elling in the Life Sciences (2014). [16] M. Bertsch, B. Franchi, N. Marcello, M. C. Tesi, and A. Tosin, Mathematical Medicine and Biology p. dqw003 (2016). [17] P. C. Bressloff and J. M. Newby, Reviews of Modern Physics 85, 135 (2013). [18] E. Kuhl, J. Mech. Behavior Biomed. Mat. 29, 529 (2014). [19] A. Goriely, The Mathematics and Mechanics of Biological Growth (Springer Verlag, New York, 2017). [20] C. R. Jack and D. M. Holtzman, Neuron 80, 1347 (2013). [21] L. A. Mihai, L. Chin, P. A. Janmey, and A. Goriely, Jour- nal of The Royal Society Interface 12, 20150486 (2015). [22] P. M. Thompson, K. M. Hayashi, G. De Zubicaray, A. L. Janke, S. E. Rose, J. Semple, D. Herman, M. S. Hong, S. S. Dittmer, D. M. Doddrell, et al., Journal of neuro- science 23, 994 (2003). [23] H. Braak and E. Braak, Acta neuropathologica 82, 239 (1991). [24] C. R. Jack Jr, D. S. Knopman, W. J. Jagust, R. C. Pe- tersen, M. W. Weiner, P. S. Aisen, L. M. Shaw, P. Ve- muri, H. J. Wiste, S. D. Weigand, et al., The Lancet Neurology 12, 207 (2013). [25] H. van den Bedem and E. Kuhl, Current Opinion in Biomedical Engineering 1, 23 (2017). [26] S. Leh´ericy, M. Marjanska, L. Mesrob, M. Sarazin, and S. Kinkingnehun, European radiology 17, 347 (2007). [27] M. Zamparo, A. Trovato, and A. Maritan, Physical re- view letters 105, 108102 (2010). [28] T. C. Michaels, S. I. Cohen, M. Vendruscolo, C. M. Dob- son, and T. P. Knowles, Physical review letters 116, 038101 (2016). [29] A. Lloret-Villas, T. Varusai, N. Juty, C. Laibe, FIG. 5: Atrophy. Top row: magnetic resonance images show- ing hippocampal atrophy in Alzheimers disease [26]: yearly examination of the same subject is shown. Increasing hip- pocampal and atrophy is observed together with ventricular enlargement and widening of cortical sulci. Middle row: acti- vation time based on initial seeding in the brain stem. Bottom row: shrinking based on the activation time and toxic protein concentration with a ratio ϑgray : ϑwhite = 4 : 1. in regular sub-cellular functions such as alpha-synucleins or tau proteins. Second, it has been proposed that the overall evolution of a typical biomarker follows regular sigmoid-like curves. Third, a given disease is associated with typical atrophy patterns. There have been multiple mathematical models pro- posed for neurodegenerative diseases, but most of them focus on biochemical pathways, cellular interactions, and the formation of amyloids [27–29]. Our approach here is radically different as we study the problem at the largest possible scale of the brain. Our minimal model takes into account the main microscopic features of these diseases, but incorporated the full brain geometry and axonal directions from diffusion tensor imaging. Despite the complexity of these diseases our model recovers the three aforementioned key features of neurodegenerative disease progression and suggests that brain geometry, transport anisotropy, and mechanics play a central role in neurodegeneration and atrophy pattern formation. This work was supported by the NSF Grant CMMI 1727268 to Ellen Kuhl. The support for Alain Goriely by the Engineering and Physical Sciences Research Council of Great Britain under research grant EP/R020205/1 is gratefully acknowledged. [1] M. Jucker and L. C. Walker, Nature 501, 45 (2013). [2] J. Brettschneider, K. Del Tredici, V. M.-Y. Lee, and J. Q. Trojanowski, Nature Reviews Neuroscience 16, 109 (2015). year 0year 1year 2year 3 N. Le Nov`Ere, H. Hermjakob, and V. Chelliah, CPT: pharmacometrics & systems pharmacology 6, 73 (2017). 6